Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Flexible and stretchable inorganic optoelectronics

Open Access Open Access

Abstract

Flexible and stretchable optoelectronics have attracted attention in recent years due to their remarkable capabilities for use in wearable computers, personal monitors, and other systems. These devices are conformal to human skin or tissue due to mechanical compliance, which is extremely suitable for biomedical or clinical applications, such as bionic devices, monitors, or curing diseases. This paper presents an overview of the flexible and stretchable inorganic optoelectronics, including mechanical design, photonic analysis, fabrication processing, some examples of the devices, and their applications in biomedical engineering. First, the recent technological advances in flexible designs and fabrication strategies are summarized. Then, the flexible and stretchable devices with different functionalities are described, including light-emitting devices, photodetectors, and other optical devices. Finally, we show the applications of flexible and stretchable optoelectronics on biomedical engineering. Also, we discuss the prospects of flexible and stretchable optoelectronic devices in the future.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Optoelectronic devices are electronic devices or systems that source [15], detect [69] and control [1012] light, usually considered a sub-field of Photonics [13]. These devices are electrical-to-optical or optical-to-electrical, which include light-emitting devices (light-emitting diode (LED), laser diodes) [1417], photoelectric devices (photodiodes (including solar cells), phototransistors and photomultipliers) [1824], and photoconductive devices (photoresistors, photoconductive camera tubes) [2528]. Optoelectronic devices or systems have been enthusiastically explored for applications such as optical fiber communications [2932], precision measuring instruments [3335], health care [3638], or power source [3942]. In this century, the researches on optoelectronics focus on improving the performance as well as changing the mechanical properties. The remarkable works have been brought to the flexible and stretchable optoelectronics to research for the compatibility of biomedical systems [4346]. The researchers have reported advancing designs or fabricating skills for flexible optoelectronics that offer not only principles on the mechanical performance of semiconductors, but also new opportunities for applications on clinics or biomedical experiments [4749].

Traditionally, inorganic semiconductors [5052] are premium materials for producing high-performance optoelectronics [5358]. These inorganic semiconductor materials with active layer structures must grow on intrinsic wafers, which are rigid, bulky, hard, and stable in high temperature [5961]. For example, the fabrication of many LEDs based on AlGaInP structure starts from epitaxial growth on gallium arsenide (GaAs) wafer [62,63]. On the other side, biomedical systems are typically deformable when suffered external stress such as squeezing, bending or stretching. The mechanical properties are totally mismatched between the conventional inorganic devices and biomedical system, which limits the optoelectronic applications in bio-related fields. As early as 1992, Gustafsson et al. fabricated a fully flexible LED using poly (ethylene terephthalate) as the substrate, soluble polyaniline as the hole-injecting electrode, a substituted poly(1,4-phenylene-vinylene) as the electroluminescent layer and calcium as the electron-injecting top contract [64]. However, the challenges for fabricating flexible and stretchable inorganic semiconductors are still unsolved. The epaxial growth of functional layers will relate to high temperature, which is unacceptable for the soft substrate made from organic materials. This limitation has been solved by the application of transfer printing technologies [6570], which success to integrate the inorganic semiconductors to flexible organic substrates. The stretchability of semiconductors can be realized by stretchable structure, such as serpentine shape [7173] and island-bridge structure [7476]. The enhancement from mechanics, fabrication have improved the performance of flexible electronics. We list some papers reported on the developments of designs, fabricating techniques, devices, and applications of flexible and stretchable semiconductors, as shown in Fig. 1 [77100]. In 2011, Kim et al. have introduced epidermal electronics incorporating electrophysiological, temperature, and strain sensors, as well as transistors, light-emitting diodes, photodetectors, radio frequency inductors, capacitors, oscillators, and rectifying diodes. All these components are configured together into “skin-like” membranes that are invisible to the skin, much like a temporary transfer tattoo [78]. In 2012, Carlson et al. summarized transfer printing techniques, ranging from the mechanics and materials aspects that govern their operation to engineering features of their use in systems with varying levels of complexity [77]. In 2016, Choi et al. summarized the soft, flexible and stretchable electronics/optoelectronics being employed in biomedicine for implantable, minimally invasive, and wearable applications [7981]. The Applications include monitoring of electro-physiological signals from the skin, tissues, and organs. In 2017, Liu et al. reviewed flexible and stretchable electronics for wearable health monitoring [8287]. The integrated skin-like devices were introduced in detail and exhibited some examples of biomedical applications. In 2018, Choi et al. summarized the developments of flexible quantum dot light-emitting diodes, and showed representative examples of future interactive displays with flexible form factors, including smart glasses and/or smart lens, LEDs woven into fabric and cloth for wearable displays, ultrathin displays attached to the human skin in the form of electronic tattoos, and bendable displays utilized as foldable tablets [88]. In 2019, Kim et al. introduced the works of wearable biosensors for healthcare monitoring. The representative examples of wearable biosensors are listed, such as eyeglass-based sweat sensor, mouth-based biosensor, Graphene-based wireless bacteria sensor, integrated wearable sensor arrays, wearable chemical-electrophysiological hybrid biosensor, and smart contact-lens biosensing platform [89100].

 figure: Fig. 1.

Fig. 1. Reviews of flexible and stretchable electronics, including flexible LEDs [64]; epidermal electronics [78]; transfer printing technologies [77]; flexible and stretchable bio-electronic devices integrated with nanomaterials [7981]; Copyright 2016, Wiley, Copyright 2008, 2014, Springer Nature; flexible and stretchable electronics [8287]; flexible and stretchable displays [88]; Flexible Electronics; wearable biosensors for healthcare monitoring [89100]. Figures reused with permission from the corresponding publisher.

Download Full Size | PDF

In this paper, we review the rapid developments of the flexible and stretchable inorganic optoelectronics [101103]. We first introduce the mechanical and photonic analysis for designing high performance, flexible and stretchable inorganic semiconductors. Then advances in different transfer printing techniques are examined for efficiently fabricating flexible and stretchable optoelectronics. Examples of these optoelectronic devices are listed as different functionalities, materials, and structures. The applications in biomedical engineering of the flexible and stretchable optoelectronics are subsequently reviewed. Finally, we conclude with a discussion of the flexible and stretchable optoelectronics on the emerging trends, possible developments, and opportunities in this field.

2. Mechanical designs and photonic analysis for flexibility and stretchability

2.1 Mechanical designs

The devices based on inorganic materials are generally fabricated on rigid semiconductor wafers. The substrates serve as the foundation for the manufacturing process, such as impurity doping, spin coating, and annealing. Generally, these substrates are useless for improving the performance of the semiconductors. Moreover, some electronics need to reduce the substrates to improve the performance or optimize the heat dissipation. Also, the substrates limit the flexibility of the inorganic semiconductors. The relationship between bending shiftiness and thickness of inorganic semiconductors can be expressed as follows.

$${EI = \frac{{Eb{h^3}}}{{12}}}$$
Where EI, E, b, h represents bending shiftiness, Young’s modulus, width and thickness of the semiconductors, respectively. This equation expresses the mechanical characters of rectangle shape semiconductors, which is the most common structure of the inorganic semiconductors. The EI is proportional to the width and the cube of thickness. This property provides a route for designing flexible inorganic semiconductors: the inorganic semiconductors such as silicon (Si) wafers are brittle and rigid, but any material can be flexible if they are thin enough.

In the past several years, some functional inorganic semiconductors have been explored in applications of flexible electronics. The most successful method for realizing flexible semiconductors is the introduction of low-dimensional inorganic semiconductors, including materials with ultra-thin films (e.g., two-dimensional layered materials), one-dimensional inorganic structures (e.g., NWs, nanotubes, nanobelts), and even 0-dimensional nanostructures (quantum dots). Figure 2(a) shows typical flexible ribbons from a Si (111) wafer. The total thickness of the ribbons is less than 1 µm [103].

 figure: Fig. 2.

Fig. 2. Mechanical designs of the flexible and stretchable inorganic semiconductors. (a) SEM image of the flexible single crystal ribbon, which thickness is less than 1 µm [105]. (b) Scheme for explaining neutral layer, and finite element simulated through-thickness [106]. (c) Schematic illustration of the process for fabricating wavy ribbons on a plastic substrate [107]. (d) SEM image of GaAs wavy ribbons on PDMS substrate [108]. (e) Some examples of wavy ribbons based on different materials. From top to bottom: Silicon, AlN, PZT, respectively [111111]. (f) Fractal geometries as general layouts for stretchable electronics and FEM images of each structure under elastic tensile strain and the demonstration of Si NMs patterned into Peano layouts then bonded to a 40% pre-strained elastomeric substrate [112]. Figures reused with permission from the corresponding publisher.

Download Full Size | PDF

Complete flexible electronics include a plastic substrate, functional layers (ultra-thin semiconductors), and packaging layer, which will form a multi-layer structure eventually, as shown in Fig. 2(b) [104]. The bending strain on each layer related to the position along thickness z. Generally, a strategy called “neutral plane” is adopted for minimizing the stain loaded on the layer to protect the functional semiconductors [111]. Theoretically, the strain is zero on the neutral plane. The position of the neutral layer in a multi-layer structure can be calculated by

$${\textrm{z}({\varepsilon = 0} )= \mathop \sum \limits_{i = 1}^n {E_I}{d_i}\left[ {\left( {\mathop \sum \limits_{j = 1}^i {d_j}} \right) - \frac{{{d_i}}}{2}} \right]\left /{\mathop \sum \limits_{i = 1}^n {E_i}{d_i}}\right.}$$
where z is the distance along the z-axis in Fig. 2(b), Ei and di are the modulus and thickness of the ith layer. Base on the Eq. (2), we can significantly reduce the stain in semiconductors by setting them on the neutral plane. This theory has successfully demonstrated and applied in flexible electronic designs for improving bending performance.

Stretchable semiconductors can be achieved by specific designs based on flexible configurations. Individual configurations can be fabricated for yielding geometries that can be stretchable. For example, we can build buckled ribbons on a stretchable substrate for realizing stretchable inorganic semiconductors. Generally, the ribbons are coupled on the polydimethylsiloxane (PDMS) surface by Van de Waals force. This force is not very strong and can be modulated by exposing the surface of PDMS with ultraviolet light (240-260 nm) before the transfer printing process. The pre-strained technology is used to form the buckled structure, as shown in Fig. 2(c) [105]. Figure 2(d) shows the buckled ribbons of GaAs on the PDMS substrate [106]. The PDMS substrate (thickness ∼4 mm) is loaded a large uniaxial pre-strain ${\varepsilon }$, and then transfer-print the GaAs ribbons on the PDMS surface. The wavy configuration will appear when we release the stain on PDMS. The height of the buckles will increase if we use larger ${\varepsilon }$. The width and length of the initial ribbons determine the dimensions of the wavy form. For example, in terms of the simple single layer of signal crystal silicon on PDMS, the wavy configuration can be calculated by

$${{\lambda _0} = \frac{{\pi h}}{{\sqrt {{\varepsilon _0}} }},\; {A_0} = h\sqrt {\frac{{{\varepsilon _{pre}}}}{{{\varepsilon _c}}} - 1} }$$
where ${{\varepsilon }_c}$, ${{\varepsilon }_{pre}}$ are the critical strain and the level of pre-strain, respectively. Here, ${{\varepsilon }_c}$ can be calculated by Poisson ratio v and Yong’s modulus E of silicon and PDMS. The wavy wavelength $({{{\lambda }_0}} )$ and amplitude and (${A_0})$ can be determined by the thickness (h) of the semiconductors [112]. The previous works have been reported a lot of inorganic ribbons based on different materials, such as Si, AlN, and ferroelectric materials, as shown in Fig. 2(e) [107109].

Although the wavy strategy provides a universal method for achieving stretchability, the stretchability of the wavy is limited by the configuration of the waveform. Generally, the extreme of applied strain on the waveform is less than 55%. The concept in fractal geometry can be exploited in stretchable electronics for improving the stretchable performance. Figure 2(f) shows some deterministic fractal designs, including line (Koch, Peano, Hilbert), loop (Moore, Vicsek) and branch-like geometries (Greek cross) [110]. The results of the finite element method (FEM) show that the uniaxial strain can be up to 75% without crack. These structures are also very suitable for the interconnections of stretchable semiconductor units. Experiments show that Si nanowires into second-order Peano layouts can be elastically strained by 105% as well as a maximum principle strain of 1% in the silicon. These results demonstrate the mechanical enhancement of the stretchable semiconductors by using the fractal geometries.

2.2 Photonic analysis

Low dimensional materials [114122] are easy to deform compared with their bulk counterparts, but the optical properties are also tuned by elastic strain [123]. For example, previous works have observed significant energetic red-shift of the near-band-edge emission in bent semiconductors. The flexoelectric effect can be used to fabricate devices such as optopiezotronics, but the stability of the electronic or optoelectronic performance will be an influence when bending. Fu et al. investigate the continuously varying electric bandstructure profile when ZnO nanowires (NWs) loaded an inhomogeneous elastic strain gradient, as shown in Fig. 3(a) [113]. The NWs with high quality are bent with a special designed four-point-bending setup, and the cathodoluminescence (CL) spectra are measured at low temperature (5.5 K). Although the ${D_0}{X_A}$ luminescence peak dominates the bending region of the wire, red-shift on the spectra can be observed and compared with the emission peak in the strain-free region by 33 meV. The results demonstrate near-band-edge (NBE) emission influenced by the strain in bent ZnO NWs.

 figure: Fig. 3.

Fig. 3. Photonic analysis for observing semiconductor properties with different stain. (a) Experiments and results of CL measurements in a bent ZnO NWs. Line-scanning CL spectra on the strain-free section “I” and from the pure bending region (section “II”-“IV”) with constant strain-gradient g = 1.20% µm−1 at 5.5 K, and the three-dimensional graph of the NBE emission spectra [115]. Copyright 2014, Wiley. (b) Optical microscope image, 3D microstructure, and profile details of a single ribbon for three samples with different waveform [116]. The right figures show the PL spectra with continuous position intervals in a single period, normalized Raman spectra of GaAs long-wavelength optical phonon on three particular positions, and periodic band gap shift in one more period of different samples. Copyright 2016, ACS.

Download Full Size | PDF

The strain-photonic coupling effect in optoelectronic materials is also obvious due to the tuning of bandgap when the semiconductors are bent [124129]. Our group (Wang et al.) reveal the continuous strain distribution in GaAs nanoribbons by applying structural buckling, as shown in Fig. 3(b) [114]. The general bending experiments can only test the optoelectronic properties (e.g. µ-Raman spectra) at a specific radius. The wavy geometry provides a route to measure continuous photoluminescence due to the different radius on the wavy. The Raman spectra show a clear bandgap shift at different location of the periodic GaAs ribbons. In the experiments, the bandgap shifts versus per unit strain are calculated by 12.8, 12.9, and 10.9 meV/% for three wavy samples with different waveform. A change rate (∼12 meV/%) is identified to modify the <001> GaAs band structure with actual stain applied on ultra-thin films. These works reveal the relationship between optoelectronic performance and strain, which are essential for the design of flexible and stretchable semiconductors in the future.

3. Fabrication strategies and transfer printing techniques

The substrate materials of flexible and stretchable electronics are generally made by organic materials, such as silicone elastomers [130133], polyimide [134136], and other plastic [137142]. These materials offer high stretchability as well as excellent adhesion for loading the inorganic semiconductors. However, the specific steps for fabricating inorganic electronics relates to high temperature, which exceeds the glass-transition and/or thermal decomposition temperatures of the plastic substrate. Transfer printing techniques [143150] are proposed for avoiding this challenge during the fabricating process of flexible and stretchable electronics.

Figure 4(a) shows a schematic illustration of the generic process flow for transfer printing solid objects, which is realized by Meitl et al [151]. and successes to transfer printing ultra-thin inorganic semiconductors from donor substrate to receiver substrate. The process begins with the preparation of a donor substrate. The functional layer can be epitaxial growth on the donor substrate by the traditional semiconductor technologies. The essential tool used here is transfer printing stamp. Generally, the stamps are made by polymeric materials. The original materials for stamps are PDMS with a smooth surface, and there are other materials used for stamps with the developments of transfer printing technology. The soft elastomeric stamps can pick up the ultrathin semiconductors by generalized adhesion forces, and most of the forces are Van der Waals interaction. The adhesion forces can be controlled by surface treatment or kinetical sensitive due to the viscoelastic behavior of the elastomer. For example, the high peel velocity (e.g., ∼10 cm/s or fast) results in strong forces to peel off the semiconductors (also called “inks”) from donor substrate, whereas low peel velocity (∼1 mm/s) is helpful to print the inks on the receiver substrate. Figure 4(b) shows a critical energy release rates (G) for the stamp/inks and semiconductor/inks interfaces [152]. The characteristic energy release rate G is related to the peel velocity. In conclusion, there are three key elements involved in a transfer printing process: substrate, ink, and stamp.

 figure: Fig. 4.

Fig. 4. Transfer printing technologies. (a) Schematic illustration of the generic process flow for a typical transfer printing process using PDMS stamp [151]. (b) Schematic diagram of critical energy release rates for the ink/substrate interface and the stamp/ink interface. The intersection of the horizontal line in the middle with the curve represents the critical peel velocity for the kinetically controlled transfer printing. The horizontal lines at the bottom and top represent very weak and strong film/substrate interfaces, respectively, corresponding to pick up only and printing only [152]. (c) Schematic illustration of the surface-treatment transfer printing process to pattern the PDMS substrate in the subtractive and addictive switchable transfer modes. The PDMS can be patterned to different shapes, as shown in right optical images [153]. (d) Octopus-inspired pads for the transfer printing process. The adhesive force is reversibly switched by the hydrogel actuation in response to the environmental temperature. The smart stamps can transfer print semiconductors with different materials, such as Si ribbons, InGaAs ribbons, or heterostructure [154]. (e) Programmable transfer printing process via automated laser writing on a micropatterned shape memory polymer stamp. The silicon and GaAs nanoribbons can be hybrid printed on PDMS, and also printed patterns of a cross [155]. Figures reused with permission from the corresponding publisher.

Download Full Size | PDF

From the above analysis, the modulation of G value is critical to complete a transfer printing process. Here we mark the separation at either the ink/donor stamp/ink and ink/receiver interface as $G_c^{ink/donor}$, $G_c^{stamp/ink}$ and $G_c^{ink/receiver}$. The relationship of G at different interference must satisfy the following equation.

$${G_c^{\frac{{stamp}}{{ink}}} > G_c^{\frac{{ink}}{{donor}}}}$$
$${G_c^{\frac{{stamp}}{{ink}}} < G_c^{\frac{{ink}}{{receiver}}}}$$
The modulation of G can be achieved by other means such as surface treatment. Wang et al. reported the switchable transfer printing (sTP) technique by the surface treatment-assisted method. With the surface treatment with different receipt of oxygen plasma (OP) and ultraviolet ozone (UVO), they realize PDMS decal inks transferred from the “soft side” to the “hard side”, and irrespective of the “soft” side coming from the PDMS stamp or the target PDMS substrate. These works demonstrate the transfer printing can be realized even the substrates and stamps made by the same materials. In their works, a PDMS stamp with square arrays of micro-posts (5 mm in diameter and 1.8 mm in height with a center-to-center separation of 10 mm), is used for the switchable transfer. The soft adjustment can be controlled by the surface oxidation. For example, OP and UVO 30 min can achieve “soft” and “hard” PDMS surface respectively. Finally, A switchable transfer is conducted on ultra-thin PDMS films (∼70 µm thick), and various patterned thin PDMS substrates are fabricated using stamps after surface oxidation, as shown in Fig. 4(c) [153].

Lee et al. provided another technique to adjust the G value to realize switchable adhesion. They demonstrate smart adhesive stamps by mimicking the muscle actuation to control pressure-induced adhesion of octopus suckers, as shown in Fig. 4(d) [154]. This mimic pad is fabricated with a hole patterned PDMS. The adhesion of the stamp pads is controlled by the thermal stimulus. When the pads stay in low temperature, the internal cavity will be small. The cavity volume will be sharply changed at a higher temperature. With the increase of the cavity volume, the pressure in the cavity is larger than the external environment value. Then the ultra-thin semiconductor will be picked up, just as resembling the actuating muscle in an octopus sucker for the reversible adhesion activity. This adhesive pad exhibits an adhesive strength of 94 kPa in response to the temperature change between 22 and 61 °C and can be used to transfer printing semiconductors such as Si and InGaAs micro-ribbon.

The easily controlled adhesion switchability is highly desired for ideal transfer printing. In previous works, our group proposed laser-induced forward transfer (LIFT) with shape memory polymer (SMP) for printing solid pixels with high resolution, as shown in Fig. 4(e) [155]. The micropatterned SMP stamps are fabricated as cone arrays. When the SMP stamp is heated above its glass transition temperature, the SMP will become soft, and the cone deformed when pressing, and the cone can be used to pick up target semiconductors. Then the cone shape is unchanged even at the room temperature. The laser can heat the cone up to glass transition temperature again, and the cone will release the semiconductors to the receiver substrate because the cone deformation decreases the contact surface with semiconductors. The transfer process can be programed by controlling the laser position and intensity by computer. It can be used for hybrid transfer printing semiconductors with different materials to a soft substrate, which provides a route for scalable integration of sophisticated electronic devices.

4. Flexible and stretchable optoelectronic devices

4.1 Light-emitting devices

Light-emitting elements are the most common and useful optoelectronic devices for displaying information. The flexible light-emitting devices include inorganic LEDs [160162], organic LEDs (OLEDs) [163167], plastic LEDs (PLEDs) [64,168170] and quantum dots (QDs) [171,172]. Although the thin-films of organics obtained many promising results of flexible screens or foldable displays, the low carrier mobilities and unstable performance are still limitations for their developments. On the contrary, traditional inorganic semiconductors exhibit excellent stability and high carrier mobility. Many research works focus on bright, inorganic electroluminescent devices, such as µLEDs, arrays, and displays. Park et al. have fabricated stretchable microscale inorganic light-emitting diodes by wavy configurations [156]. They assemble small inorganic LEDs into addressable arrays on plastic substrates. The epitaxial semiconductor layers include AlInGaP quantum well structures (6-nm-thick In0.56Ga0.44P wells, with 6-nm-thick barriers of Al0.25Ga0.25In0.5P on top and bottom), cladding films (200-nm thick layers of In0.5Al0.5P:Zn and In0.5Al0.5P:Si for the p and n sides, respectively), spreaders (800-nm-thick layers of Al0.45Ga0.55As:C and Al0.45Ga0.55As:Si for the p and n sides, respectively), and contacts (5-nm-thick layer of GaAs:C and 500-nm-thick layer of GaAs:Si for the p and n sides, respectively), for a total thickness of ∼2.523 mm, all grown on AlAs (1500-nm-thick layer of Al0.96Ga0.04As:Si) on a GaAs substrate. Then the functional layer is moved from a GaAs substrate to PDMS substrate. The passive matrix, stretchable inorganic LED display that uses a noncoplanar mesh configuration, on a rubber substrate is shown in Fig. 5(a). Also, the flexible LEDs array with different light wavelength can be achieved by transfer printing methods. Different from red LEDs, the blue LEDs are fabricated by gallium nitride-based (InGaN) materials, as shown in Fig, 5(b) [157]. The InGaN epitaxial stacks are grown on Si wafers with (111) orientation. The release of these LEDs from silicon substrate is utilized by a wet chemical etching process with potassium hydroxide (KOH) or tetramethylammonium hydroxide. The flexible blue light can generate white light by integrating with phosphors on PDMS. The color will be different with different thickness of phosphors, and the chromaticity follows an approximately linear path between blue and yellow, which is determined by both the blue emission of LEDs and yellow radiation of phosphor.

 figure: Fig. 5.

Fig. 5. Flexible and stretchable light-emitting devices. SEMs and optical images of the red (a) and (b) blue LED array. The red LEDs are based on AlInGaP structures (50 mm by 50 mm) created by vertical, patterned etching an epitaxial multilayer stack grown on a GaAs wafer [156]. The blue LEDs are InGaN µ-ILED arrays grown on Si (111) wafer [157]. Both of them are moved to a plastic substrate by transfer printing techniques. (c) Wearable micro light-emitting diodes integrated on a lab coat [158]. (d) RGB displays with transfer-printed assemblies of 8 µm × 15 µm inorganic light-emitting diodes [159]. (e) A stretchable high-density array of blue LEDs. There are 500*500 units connected by stretchable wire. Figures reused with permission from the corresponding publisher.

Download Full Size | PDF

One of the essential applications of flexible or stretchable light-emitting devices is for wearable displays. Wearable displays are a potential candidate for future information exhibitions. Lee et al. have reported wearable micro LEDs (WµLEDs) integrated with wireless power supplement [158]. The 30*30 WµLED arrays are transferred by transparent elastomeric adhesive from GaAs substrate and combined with a wireless energy transfer system fabricated on a 100% cotton fabric. The strain of WµLEDs can be reached 100%, and the array showed excellent stability under 85 °C and 85% relative humidity. This array can be stitched onto a lab coat, which demonstrates its practical applications, as shown in Fig. 5(c).

Bower et al. fabricated emissive displays by transfer printing 8 µm *15 µm inorganic light-emitting diodes [159]. The displays integrate red light ((Al)InGaP quantum wells, and were grown on GaAs substrates), green and blue µILEDs growth taking place on a Si (111) wafer. The transfer stamp is made from glass and elastomer by injection-moulding technique. The shapes of the stamp are specific design for transfer printing LEDs with a different color. The total thickness of plastic µILED display (100 × 100, 254 PPI) is ∼130 µm, as shown in Fig. 5(d).

Our group has fabricated a 500*500 units array of blue LEDs. These blue LEDs are InGaN epitaxial structure grown on a silicon substrate. Then the functional layer is transfer printed on PDMS, and the silicon substrate is etched by oxygen plasma. The LED unit is a square with a side length of 150 µm, and the spacing between the units is 10 µm. This LEDs array can emit white light with the and yellow radiation of phosphor, and the high density of LEDs makes a high resolution.

Stretchable inorganic light-emitting arrays can be worn on the location of the body that highly deformable, which provide a possibility of tattoo-like or epidermal displays [175]. These displays are designed as stretchable structures due to the use of fractal shape and island-bridge structures. Choi et al. fabricated a stretchable, active matrix (AM) inorganic LED display with rapid response time and low power consumption [173]. Most of the applied strain is about 62%, which is limited by the serpentine bridges, and the images of stretchable displays are shown in Fig. 6(a).

 figure: Fig. 6.

Fig. 6. Mechanical designs for the stretchable LED array. (a) Stretchable active-matrix inorganic light-emitting diode display. These images show uniform emission characteristics of the stretchable LED display under uniaxial strain from 0% to 40%; white lines indicate the horizontal direction [173]. (b) Stretchable LED chain with strain isolation design. The substrate can be spatio-programmed rigidity, which means the substrate is soft and stretchable expect the section of LEDs [174]. Figures reused with permission from the corresponding publisher.

Download Full Size | PDF

Despite the rapid developments of stretchable optoelectronics for human-machine interface, the robustness is still an unresolved challenge for commercial use of these stretchable displays. Our group works have proposed an approach that uses a new polymer substrate with programmable rigidity to realizing strain isolation, as shown in Fig. 6(b) [174]. The plastic substrate used here is spatio-programmed substrate rigidity, which Yong’s modulus can be changed by H2O2. A LED chain is fabricated for demonstrating the strain isolation. The substrate under LEDs is rigid status, while the other parts are stretchable. The LEDs remain operational by uniaxial strain (21% global strain, corresponding to 36% strain at the interconnects.

4.2 Photodetectors

Photodetectors, also called photosensors, are able to convert light intensity into electrical signals, which are indispensable components for the light sensor. They have been applied in a variety of emerging areas such as wearable, implantable, or printed optical devices for light detection or perception. Previous works reported various flexible and stretchable photodetectors based on inorganic materials. Here, we introduce some examples of inorganic or inorganic-organic hybrid photodetectors on a flexible and stretchable substrate.

Silicon photodetectors can be fabricated on Si-based complementary with low-cost technologies, which is attracting growing interest in recent years. Seo et al. report flexible phototransistors on single-crystalline Si nanomembrane (Si NM), as shown in Fig. 7(a) [176]. The silicon on insulation (SOI) wafer with 270 nm p-typed doped top Si layer and a 200 nm buried oxide layer. The top Si can be used for fabricating ultra-thin photoresistors, and the oxide layer acts as a sacrificial layer for transfer printing. The device exhibits stable responsivity with less than 5% variation during bending at small radii of curvatures (up to 15 mm). GaAs is another relevant material for yielding photodetectors. Lee et al. report an array of ultrathin GaAs solar microcell grown epitaxially on GaAs wafers, and then transferred on elastomeric substrates, as shown in Fig. 7(b) [177].

 figure: Fig. 7.

Fig. 7. Flexible and stretchable inorganic/inorganic-organic hybrid photodetectors. (a) Flexible phototransistors based on transferrable single-crystalline Si nanomembrane (Si NM). The device exhibits stable responsivity with less than 5% of variation under bending at small radii of curvatures (up to 15 mm) [176]. (b) Optical images of the stretchable GaAs photovoltaic modules. The photodetector array is designed as an island-bridge structure for stretchability [177]. (c) Images and switching behaviors of an intrinsically stretchable nanowire photodetector based on ZnO NM. The ZnO NM layer is embedded in the PDMS matrix, yielding a structure with all the components (electrodes and detection channels) being intrinsically stretchable [191]. (d) Silicon-organic hybrid solar cells by the vertical array of Si micro-pillars embedded into elastomeric substrates [192]. (e) Flexibility organic-inorganic perovskite solar cells. The recoverable shape polymer is utilized as a substrate of perovskite solar cell to enable complete shape recovery of the device upon sub-millimeter bending radii [193]. Figures reused with permission from the corresponding publisher.

Download Full Size | PDF

Nanowire materials have been explored in flexible and stretchable applications in recent years. The advantage of one-dimensional nanostructures makes these materials suitable for flexible electronics. There are several nanowires used for flexible photodetectors, and the based materials include ZnO [178,179], SnO2 [180182], ZnGa2O4 [183185], single-wall carbon nanotubes (SWCNT) [185,186], ZnO-CdO [187189] and ZnO-Cds [189,190]. Figure 7(c) shows the performance and deforming images of ZnO nanowire photodetector [191]. These photodetectors are mainly focused on visible-blind UV wavelength sensing, which is essential to broad range applications such as skin protection and space exploration.

Inorganic–organic hybrid composites can achieve possible band structure modulation and charge trapping effect at the interface of organic/inorganic components. These characteristics are attractive to obtain high-performance photodetectors due to the efficiently facilitate charge separation and transportation. Recently, there is a rapid development of inorganic-organic hybrid technologies for flexible or stretchable photodetectors. Figure 7(d) is an example of stretchable, bifacial Si-organic hybrid solar cells fabricated on elastomeric substrates [192]. The solar cells can be stressed up to 100% without crack due to the vertical array of Si micro-pillars designs. The using of organic materials with specific mechanical designs can achieve a delightful deformation ability. The other materials can be used for organic-inorganic hybrid photodetector include P3HT [194196], Indium gallium zinc oxide (IGZO) [197199], CdSe [200,201], and others.

Perovskite is a potential material to achieve inorganic-organic hybrid photodetectors, which has already demonstrated by using it to fabricate high-performance devices such as perovskite solar cells. Park et al. focus on the optimization of the flexibility of perovskite solar cell, as shown in Fig. 7(e) [193]. The 13.5 GPa elastic modulus of the perovskite is obtained, and the stress and strain distribution in the device at small bending radii (r = 1 mm and r = 0.5 mm) is observed. Plastic deformation, as well as a significant drop in the power conversion efficiency (PCE) value, is experienced at a bending radius of 0.5 mm.

4.3 Other optical devices

The flexible and stretchable photonic devices are promising solutions for precision instruments or bandwidth interconnections. Photonic devices with different functionalities have already demonstrated on rigid substrates, such as a waveguide, filters, grating device, and sensors. In recent years, researchers have reported flexible and stretchable photonic devices that use inorganic materials. Figure 8 show examples of flexible and stretchable photonic devices. Figure 8(a) illustrates a flexible photonic device that integrates waveguides, micro-mirrors, and optoelectronic components [202]. The thin foil of 150 µm thickness with embedded active optical can be realized low-loss links for function interconnecting. A strain sensor is realized by the optical resonator. Figure 8(b) shows its undeformed and stretched states [203]. The optical performance of the devices is demonstrated with 36% nominal strain. Li et al. demonstrate single-mode stretchable integrated photonic devices, as shown in Fig. 8(c) [204]. Cho et al. report a continuously tunable color filter based on a self-assembled isotopically stretchable microbead monolayer, which can be used for measuring lateral strain, as shown in Fig. 8(d) [205].

 figure: Fig. 8.

Fig. 8. flexible and stretchable photonics for interconnection and sensing. (a) Integration photonics for signal transmission. The device integrates waveguides, micro-mirrors, and optoelectronic components [202]. (b) Flexible silicon photonic devices. An optical strain sensor uses 5 µm-radius micro ring resonators on the flexible substrate [203]. (c) Single-mode stretchable integrated photonic devices using chalcogenide glass (ChG) and epoxy polymer [204]. (d) Continuously tunable color filter based on a self-assembled isotopically stretchable microbead monolayer. The spectra of the filtered light are solely controlled by external strain (up to 32% radial strain) to cover a broad visible spectrum [205]. Figures reused with permission from the corresponding publisher.

Download Full Size | PDF

5. Biomedical applications

In this section, we will provide some examples of the biomedical applications of the flexible and stretchable inorganic optoelectronics. The flexible sensors or devices can be useful and efficient in combination with biomedical applications due to the similar mechanical properties with skin or tissue. There are researches focus on the stretchable devices imitating biomedical organs, and also some reports of physical sensors for monitoring vital signs. Ko et al. fabricated a digital camera inspired by the arthropod eye, which is comparable in number (180) to those of the eyes of fire ants (Solenopsis fugax) and bark beetle, as shown in Fig. 9(a) [206]. The thin, silicon photodiodes (active areas d2 ≈ 160 mm*160 mm) and blocking diodes consist of the matching array in an open mesh configuration with capability for matrix addressing. The patterned PDMS consists of an array of 163 16 convex microlenses (with a radius of curvature of each microlens r < 400mm). This compound eye device demonstrated the flexible and stretchable photodetectors could act as a replacement for a real eye. Zhang et al. also fabricate hemispherical electronic eye systems with origami silicon optoelectronic, as shown in Fig. 9(b) [207]. The digital image sensor reported here is fabricated by origami approach, which provides a method for fabricating three-dimensional flexible electronics.

 figure: Fig. 9.

Fig. 9. Representative flexible and stretchable inorganic optoelectronics for biomedical applications. (a) Images of components for a digital camera that takes the form of a hemispherical, apposition compound eye. The photodiode array is fabricated by silicon, and the small hemisphere lens is made by PDMS [206]. (b) Geometric origami of silicon optoelectronics for the hemispherical electronic eye. 676 polygon blocks consisting of pentagons and hexagons were mapped into a net of a subdivided half truncated icosahedron, which was then folded to form a hemisphere [207]. (c) Images and performance of the released filter membrane on a flexible PDMS sheet. This filter structure consists of a series of titanium dioxide (TiO2) and silicon dioxide (SiO2) films with a total thickness of about 8 µm. The films are biocompatible, and no apparent signs of pathological inflammatory tissue responses to the filter implantation are observed after 5 weeks [208]. (d) Injectable optoelectronics with applications for wireless optogenetics. A custom flexible polyimide film–based lightweight (∼0.7 g) power scavenger or a rigid printed circuit board-based scavenger (∼2.0 g) can be temporarily mounted on freely moving animals for short-term experimentation without constraint in natural animal behavior [209]. Figures reused with permission from the corresponding publisher.

Download Full Size | PDF

Liu et al. report thin-film optical filters that can be used for biomedical applications due to the excellent vivo compatibility, as shown in Fig. 9(c) [208]. The filter can reflect blue and green photons (420–560 nm) and transmits yellow and red photons (560–620 nm). The materials for fabricating the filters include titanium dioxide (TiO2) and silicon dioxide (SiO2) films with a total thickness of about 8 µm. The package of PDMS can improve the biocompatibility, which is demonstrated by the vivo experiments. These filter membranes can combine with optoelectronic devices for further applications.

Flexible and stretchable optoelectronics are also suitable for yielding injectable or implantable devices due to similar mechanical properties with biomedical systems. Ultra-thin semiconductors with high performance are helpful for scientific research of clinics. Kim et al. introduce injectable class optoelectronics that can insert light sources, detectors, sensors, and other components into precise locations of the deep brain for optogenetic research. The gallium nitride (GaN) µ-ILED (6.45 µm thick, 50 × 50 µm2) is integrated on a flexible substrate, and the total thickness of the injected multifunctional optoelectronic systems is only ∼20 µm. This exceptionally thin geometry, low bending rigidity, and a high degree of mechanical allow for minimally invasive operation, as shown in Fig. 9(d).[209]

Flexible and stretchable optoelectronic offer a wide range of unprecedented opportunities for monitoring vital signs. The ability of synchronous deformation with skin can achieve long-term monitoring without irritation to human skin or tissue. The comprehensive capabilities of bending, stretching or twisting also provide outstanding conformability to skin. The flexible optoelectronics can monitoring pulse rate and oxygen blood saturation, which is the critical indicators for evaluating healthy [213217]. Although there are many commercial oximeters, most of them are fabricated on traditional printed circuit board (PCB) and integrated on a smartwatch or wrist band. The stress or squeeze is loaded on the skin when wearing these devices. Also, these oximeters can work at a specific location of the body, which limits the applications of healthcare monitoring in terms of different scenarios. Flexible or stretchable optoelectronics can solve the problems mentioned above due to the deformable and imperceptible characteristics. Kim et al. report miniaturized battery-free wireless systems using inorganic semiconductors, as shown in Fig. 10(a) [210]. This reflectance oximeter using red infrared LEDs and A photodetector positioned in between these two LEDs captures the backscattered light from the blood. These measurements are possible, mostly independent of variations in the shape, thickness, and optical properties of the nail, due to sufficient signal to noise ratio and intimate contact. Operation on the fingernail is highly desirable for long-term operation, with minimized motion artifacts and risks for irritation.

 figure: Fig. 10.

Fig. 10. Flexible and stretchable oximeters. (a) Miniaturized battery-free wireless systems for wearable pulse oximetry. The wireless transmitting components are integrated with inorganic optoelectronics, and the systems can conformal contact to the target probing area (i.e., fingernail, toenail, or various skin locations) for avoiding mechanical damage [210]. (b) The performance analysis via Monte Carlo (MC) simulation. The deformation of flexible devices will cause an unstable optical path. Mechanical optimization is needed to keep the detected signal stable during deformation [211]. (c) Epidermal optoelectronics with strain isolation design for blood oxygen monitoring. The all-in-one free-floating structure is adopted to keep the stable position of LEDs and photodetectors, which stabilize the optical signals for precise blood oxygen monitoring during deformation [212]. Figures reused with permission from the corresponding publisher.

Download Full Size | PDF

Our group publishes the optical and mechanical optimization of the epidermal optoelectronics for monitoring blood oxygen saturation. We first simulate the optical path in human tissue for analyzing the influence of deformation to detected signals, as shown in Fig. 10(b) [211]. A 2D tissue model is built, and the distribution of exited light from human tissue is investigated by Monte Carlo (MC) simulation. The performance with different distance between LEDs and photodetector, light incident angle and intensity is quantified by simulation. Then a systematic strategy to design an epidermal inorganic optoelectronic device for blood oxygen monitoring is proposed by using specific strain-isolation design, nanodiamond thinning, and hybrid transfer printing, as shown in Fig. 10(c) [212]. The all-in-one free-floating structure can isolate the strain loaded on LEDs and photodetectors, which results in a fixed distance position between LEDs and photodetectors when stretching. The optical signals are stable when the device is deformed. These epidermal oximeters achieve not only flexibility and stretchability but also the stability of the optical path, which improves the performance of optoelectronics for biomedical application.

6. Summary and perspective

In this paper, we reviewed the latest developments of flexible and stretchable inorganic optoelectronics, including design, fabrication, devices, and the applications on biomedical applications. The combination of organic mechanical properties and inorganic functional devices provides several promising engineering options for fully integrated systems in future distributed or mixed form. The researches of stretchable devices are very complex, which involve materials growth, heterogeneous integration, geometry design and micromechanics, and adhesion and interface science. For example, the modulus of silicon is about 100,000 times that of a typical elastomer; the thermal conductivity is ∼1000 times, and the coefficient of thermal expansion is ∼100 times smaller [218]. For solving this extreme mismatch, the design of reasonable mechanical structures is required. The challenges are even more pronounced for more complex, large-scale integrated stretchable devices because the device deformation can have a massive influence on device performance. The researches on these complex systems require comprehensive consideration of materials science and effective thermal management to ensure mechanical reliability. Stretchable electronic devices must not only achieve large-scale integration but also ensure reliability in tens of thousands of stretching processes for commercial use in the future.

At present, the results of stretchable devices are changing our concept of electronics. The electronics are no longer hard and rigid but soft and stretchable as human skin. They have potential applications in many fields, and some of the most compelling applications are biomedical related devices that solve critical issues in personal health monitoring and treatment. The soft and elastic mechanical properties are more incoordinate with the actual situation of biomedical tissues. Such skin-like devices, combined with the use of biocompatible materials, significantly improve human comfort when exposed to a biomedical interface. Therefore, stretchable devices are capable of long-term monitoring tasks. At present, the main works in this area focus on integrating flexible LEDs and photodetectors and utilizing the spectral absorption characteristics of human tissues to realize vital signs monitoring. In addition to biomedical applications, stretchable electronic devices can also be used in industries such as optical communications or robotics. Even they may be used in future cellular phones, as the recently released foldable mobile phones involve the perceived form of mobile phones. The basic ideas can also be exploited in other semiconductor technologies, including electromagnetism or acoustics. These and other related engineering opportunities, together with a wide range of fascinating scientific themes, provide a powerful promotion for applications with critical social impacts.

Funding

National Basic Research Program of China (973 Program) (2015CB351900); National Natural Science Foundation of China (11320101001, 11625207).

References

1. N. W. Rosemann, J. P. Eußner, A. Beyer, S. W. Koch, K. Volz, S. Dehnen, and S. Chatterjee, “A highly efficient directional molecular white-light emitter driven by a continuous-wave laser diode,” Science 352(6291), 1301–1304 (2016). [CrossRef]  

2. J. W. Sun, J. Y. Baek, K.-H. Kim, C.-K. Moon, J.-H. Lee, S.-K. Kwon, Y.-H. Kim, and J.-J. Kim, “Thermally activated delayed fluorescence from azasiline based intramolecular charge-transfer emitter (DTPDDA) and a highly efficient blue light emitting diode,” Chem. Mater. 27(19), 6675–6681 (2015). [CrossRef]  

3. Y. J. Cho, K. S. Yook, and J. Y. Lee, “Cool and warm hybrid white organic light-emitting diode with blue delayed fluorescent emitter both as blue emitter and triplet host,” Sci. Rep. 5(1), 7859 (2015). [CrossRef]  

4. C. Lee, C. Shen, C. Cozzan, R. M. Farrell, J. S. Speck, S. Nakamura, B. S. Ooi, and S. P. DenBaars, “Gigabit-per-second white light-based visible light communication using near-ultraviolet laser diode and red-, green-, and blue-emitting phosphors,” Opt. Express 25(15), 17480–17487 (2017). [CrossRef]  

5. A. A. Alatawi, J. A. Holguin-Lerma, C. H. Kang, C. Shen, R. C. Subedi, A. M. Albadri, A. Y. Alyamani, T. K. Ng, and B. S. Ooi, “High-power blue superluminescent diode for high CRI lighting and high-speed visible light communication,” Opt. Express 26(20), 26355–26364 (2018). [CrossRef]  

6. X. Zhou, L. Gan, W. Tian, Q. Zhang, S. Jin, H. Li, Y. Bando, D. Golberg, and T. Zhai, “Ultrathin SnSe2 Flakes Grown by Chemical Vapor Deposition for High-Performance Photodetectors,” Adv. Mater. 27(48), 8035–8041 (2015). [CrossRef]  

7. G. Lou, H. Zhu, A. Chen, Y. Wu, Z. Chen, Y. Ren, Y. Liang, J. Li, X. Gui, and D. Zhong, “Single and Two-photon Absorption Single-microbelt Photodetector,” in Asia Communications and Photonics Conference, (Optical Society of America, 2017), Su2A. 167.

8. Z. Ling, R. Yang, J. Chai, S. Wang, W. Leong, Y. Tong, D. Lei, Q. Zhou, X. Gong, and D. Chi, “Large-scale two-dimensional MoS 2 photodetectors by magnetron sputtering,” Opt. Express 23(10), 13580–13586 (2015). [CrossRef]  

9. Y. Zhang, D. J. Hellebusch, N. D. Bronstein, C. Ko, D. F. Ogletree, M. Salmeron, and A. P. Alivisatos, “Ultrasensitive photodetectors exploiting electrostatic trapping and percolation transport,” Nat. Commun. 7(1), 11924 (2016). [CrossRef]  

10. Y. Yuan, D. Wang, B. Zhou, S. Feng, M. Sun, S. Zhang, W. Gao, Y. Bi, and H. Qin, “High luminous fluorescence generation using Ce: YAG transparent ceramic excited by blue laser diode,” Opt. Mater. Express 8(9), 2760–2767 (2018). [CrossRef]  

11. J. Yu, S. Jin, Q. Wei, Z. Zang, H. Lu, X. He, Y. Luo, J. Tang, J. Zhang, and Z. Chen, “Hybrid optical fiber add-drop filter based on wavelength dependent light coupling between micro/nano fiber ring and side-polished fiber,” Sci. Rep. 5(1), 7710 (2015). [CrossRef]  

12. L. De Santis, C. Antón, B. Reznychenko, N. Somaschi, G. Coppola, J. Senellart, C. Gómez, A. Lemaître, I. Sagnes, and A. G. White, “A solid-state single-photon filter,” Nat. Nanotechnol. 12(7), 663–667 (2017). [CrossRef]  

13. Wikipedia, “Optoelectronics,” (2019), retrieved https://en.wikipedia.org/wiki/Optoelectronics.

14. M. Olle and A. Viršile, “The effects of light-emitting diode lighting on greenhouse plant growth and quality,” Agric. Food Sci. 22(2), 223–234 (2013). [CrossRef]  

15. K. Lin, J. Xing, L. N. Quan, F. P. G. de Arquer, X. Gong, J. Lu, L. Xie, W. Zhao, D. Zhang, and C. Yan, “Perovskite light-emitting diodes with external quantum efficiency exceeding 20 per cent,” Nature 562(7726), 245–248 (2018). [CrossRef]  

16. Y. Liu, C. Li, Z. Ren, S. Yan, and M. R. Bryce, “All-organic thermally activated delayed fluorescence materials for organic light-emitting diodes,” Nat. Rev. Mater. 3(4), 18020 (2018). [CrossRef]  

17. X. Li, Y.-B. Zhao, F. Fan, L. Levina, M. Liu, R. Quintero-Bermudez, X. Gong, L. N. Quan, J. Fan, and Z. Yang, “Bright colloidal quantum dot light-emitting diodes enabled by efficient chlorination,” Nat. Photonics 12(3), 159–164 (2018). [CrossRef]  

18. F. Teng, K. Hu, W. Ouyang, and X. Fang, “Photoelectric Detectors Based on Inorganic p-Type Semiconductor Materials,” Adv. Mater. 30(35), 1706262 (2018). [CrossRef]  

19. W. Ouyang, F. Teng, J. H. He, and X. Fang, “Enhancing the Photoelectric Performance of Photodetectors Based on Metal Oxide Semiconductors by Charge-Carrier Engineering,” Adv. Funct. Mater. 29(9), 1807672 (2019). [CrossRef]  

20. T. Chen, X. Wang, P. Han, J.-H. Lee, C. Zhang, and Y. Zhang, “Effects of micro/nano-structures on the photoelectric properties of silicon solar cell: A pump-probe study,” in International Symposium on Ultrafast Phenomena and Terahertz Waves, (Optical Society of America, 2018), WI37.

21. K. Zhang, S. Wang, and Y. Yang, “A One-Structure-Based Piezo-Tribo-Pyro-Photoelectric Effects Coupled Nanogenerator for Simultaneously Scavenging Mechanical, Thermal, and Solar Energies,” Adv. Energy Mater. 7(6), 1601852 (2017). [CrossRef]  

22. W. S. Yang, J. H. Noh, N. J. Jeon, Y. C. Kim, S. Ryu, J. Seo, and S. I. Seok, “High-performance photovoltaic perovskite layers fabricated through intramolecular exchange,” Science 348(6240), 1234–1237 (2015). [CrossRef]  

23. M. Saliba, T. Matsui, K. Domanski, J.-Y. Seo, A. Ummadisingu, S. M. Zakeeruddin, J.-P. Correa-Baena, W. R. Tress, A. Abate, and A. Hagfeldt, “Incorporation of rubidium cations into perovskite solar cells improves photovoltaic performance,” Science 354(6309), 206–209 (2016). [CrossRef]  

24. A. Polman, M. Knight, E. C. Garnett, B. Ehrler, and W. C. Sinke, “Photovoltaic materials: Present efficiencies and future challenges,” Science 352(6283), aad4424 (2016). [CrossRef]  

25. B. Witkowski, R. Pietruszka, S. Gieraltowska, L. Wachnicki, H. Przybylinska, and M. Godlewski, “Photoresistor based on ZnO nanorods grown on a p-type silicon substrate,” Opto-Electron. Rev. 25(1), 15–18 (2017). [CrossRef]  

26. J. Liu, Y. Liang, L. Wang, B. Wang, T. Zhang, and F. Yi, “Fabrication and photosensitivity of CdS photoresistor on silica nanopillars substrate,” Mater. Sci. Semicond. Process. 56, 217–221 (2016). [CrossRef]  

27. S. Chaurasia, A. Chatterjee, S. Selvaraja, and S. Avasthi, “Infrared (IR) photoresistors based on recrystallized amorphous germanium films on silicon using liquid phase epitaxy,” in Optical Sensing and Detection V, (International Society for Optics and Photonics, 2018), 106802T.

28. J. Dai, C. Ruan, and X. Zhang, “Simulation and Analysis of Photoconductive Vacuum Diode Arrays in Terahertz Band,” in 2018 Progress in Electromagnetics Research Symposium (PIERS-Toyama), (IEEE, 2018), 1633–1636.

29. H. Venghaus and N. Grote, Fibre optic communication: key devices (Springer, 2017), Vol. 161.

30. K. Kikuchi, “Fundamentals of coherent optical fiber communications,” J. Lightwave Technol. 34(1), 157–179 (2016). [CrossRef]  

31. M. N. Sadiku, Optical and wireless communications: next generation networks (CRC press, 2018).

32. A. E. Willner, H. Huang, Y. Yan, Y. Ren, N. Ahmed, G. Xie, C. Bao, L. Li, Y. Cao, and Z. Zhao, “Optical communications using orbital angular momentum beams,” Adv. Opt. Photonics 7(1), 66–106 (2015). [CrossRef]  

33. K. Lyons, S. Pang, P. G. Kwiat, and A. N. Jordan, “Precision optical displacement measurements using biphotons,” Phys. Rev. A 93(4), 043841 (2016). [CrossRef]  

34. G. Berger and M. Wendel, “Optical Metrology of Freeforms and Complex Lenses: 3D form measurements of precision optical surfaces based on scanning point interferometry,” Opt. Photonik 13(1), 40–43 (2018). [CrossRef]  

35. H. Leopardi, J. Davila-Rodriguez, F. Quinlan, J. Olson, J. A. Sherman, S. A. Diddams, and T. M. Fortier, “Single-branch Er: fiber frequency comb for precision optical metrology with 10− 18 fractional instability,” Optica 4(8), 879–885 (2017). [CrossRef]  

36. B. Javidi, “Advances in 3D Imaging with Applications to Displays, Computational Imaging, Optical Security, and Healthcare,” in Imaging Systems and Applications, (Optical Society of America, 2016), IW5F. 3.

37. B. W. Pogue, “Biomedical Engineering or Biomedical Optics: Will the Real Discipline Please Stand Up?” J. Biomed. Opt. 24(04), 1 (2019). [CrossRef]  

38. J. Zhang, Q. Yang, K. Saito, K. Nozato, D. R. Williams, and E. A. Rossi, “An adaptive optics imaging system designed for clinical use,” Biomed. Opt. Express 6(6), 2120–2137 (2015). [CrossRef]  

39. D. H. Cao, C. C. Stoumpos, O. K. Farha, J. T. Hupp, and M. G. Kanatzidis, “2D homologous perovskites as light-absorbing materials for solar cell applications,” J. Am. Chem. Soc. 137(24), 7843–7850 (2015). [CrossRef]  

40. J. Liu, S. Chen, D. Qian, B. Gautam, G. Yang, J. Zhao, J. Bergqvist, F. Zhang, W. Ma, and H. Ade, “Fast charge separation in a non-fullerene organic solar cell with a small driving force,” Nat. Energy 1(7), 16089 (2016). [CrossRef]  

41. J.-H. Im, J. Luo, M. Franckevičius, N. Pellet, P. Gao, T. Moehl, S. M. Zakeeruddin, M. K. Nazeeruddin, M. Grätzel, and N.-G. Park, “Nanowire perovskite solar cell,” Nano Lett. 15(3), 2120–2126 (2015). [CrossRef]  

42. X. Lan, O. Voznyy, A. Kiani, F. P. García de Arquer, A. S. Abbas, G. H. Kim, M. Liu, Z. Yang, G. Walters, and J. Xu, “Passivation using molecular halides increases quantum dot solar cell performance,” Adv. Mater. 28(2), 299–304 (2016). [CrossRef]  

43. Q. Chen, M. Lin, Y. Fang, Z. Wang, Y. Yang, J. Xu, Y. Cai, and R. Huang, “Integration of biocompatible organic resistive memory and photoresistor for wearable image sensing application,” Sci. China Inf. Sci. 61(6), 060411 (2018). [CrossRef]  

44. M. J. Tan, C. Owh, P. L. Chee, A. K. K. Kyaw, D. Kai, and X. J. Loh, “Biodegradable electronics: cornerstone for sustainable electronics and transient applications,” J. Mater. Chem. C 4(24), 5531–5558 (2016). [CrossRef]  

45. R. Li, L. Wang, D. Kong, and L. Yin, “Recent progress on biodegradable materials and transient electronics,” Bioactive Materials 3(3), 322–333 (2018). [CrossRef]  

46. M. Zhu, C. Jia, Y. Wang, Z. Fang, J. Dai, L. Xu, D. Huang, J. Wu, Y. Li, and J. Song, “Isotropic paper directly from anisotropic wood: top-down green transparent substrate toward biodegradable electronics,” ACS Appl. Mater. Interfaces 10(34), 28566–28571 (2018). [CrossRef]  

47. R. Feiner and T. Dvir, “Tissue–electronics interfaces: From implantable devices to engineered tissues,” Nat. Rev. Mater. 3(1), 17076 (2018). [CrossRef]  

48. A. Tzur-Balter, G. Shtenberg, and E. Segal, “Porous silicon for cancer therapy: from fundamental research to the clinic,” Rev. Chem. Eng. 31(3), 193–207 (2015). [CrossRef]  

49. Y. R. Jeong, G. Lee, H. Park, and J. S. Ha, “Stretchable, Skin-Attachable Electronics with Integrated Energy Storage Devices for Biosignal Monitoring,” Acc. Chem. Res. 52(1), 91–99 (2019). [CrossRef]  

50. U. K. Mishra, L. Shen, T. E. Kazior, and W. Yi-Feng, “GaN-Based RF Power Devices and Amplifiers,” Proc. IEEE 96(2), 287–305 (2008). [CrossRef]  

51. Z. Chen, Y. Zhang, H. Zhang, Y. Yu, X. Song, H. Zhang, M. Cao, Y. Che, L. Jin, and Y. Li, “Low-voltage all-inorganic perovskite quantum dot transistor memory,” Appl. Phys. Lett. 112(21), 212101 (2018). [CrossRef]  

52. S. H. Lee, J. W. Kim, T. I. Lee, and J. M. Myoung, “Inorganic Nano Light-Emitting Transistor: p-Type Porous Silicon Nanowire/n-Type ZnO Nanofilm,” Small 12(31), 4222–4228 (2016). [CrossRef]  

53. C. D. Vazquez-Colon, D. C. Look, E. Heller, J. S. Cetnar, and A. A. Ayon, “Simple ohmic contact formation in HEMT structures: application to AlGaN/GaN,” in Gallium Nitride Materials and Devices XIV, (International Society for Optics and Photonics, 2019), 1091819.

54. S. Chung, P. Srivastava, X. Yang, T. Palacios, and H. Lee, “High-Performance GaN HEMT Track-and-Hold Sampling Circuits with Digital Post-Correction,” in Research Abstracts, 2018), 7.

55. P.-C. Chou, S.-H. Chen, T.-E. Hsieh, S. Cheng, J. del Alamo, and E. Chang, “Evaluation and reliability assessment of GaN-on-Si MIS-HEMT for power switching applications,” Energies 10(2), 233 (2017). [CrossRef]  

56. I. Åberg, G. Vescovi, D. Asoli, U. Naseem, J. P. Gilboy, C. Sundvall, A. Dahlgren, K. E. Svensson, N. Anttu, and M. T. Björk, “A GaAs nanowire array solar cell with 15.3% efficiency at 1 sun,” IEEE J. Photovoltaics 6(1), 185–190 (2016). [CrossRef]  

57. D. Huber, M. Reindl, Y. Huo, H. Huang, J. S. Wildmann, O. G. Schmidt, A. Rastelli, and R. Trotta, “Highly indistinguishable and strongly entangled photons from symmetric GaAs quantum dots,” Nat. Commun. 8(1), 15506 (2017). [CrossRef]  

58. G. J. Brady, A. J. Way, N. S. Safron, H. T. Evensen, P. Gopalan, and M. S. Arnold, “Quasi-ballistic carbon nanotube array transistors with current density exceeding Si and GaAs,” Sci. Adv. 2(9), e1601240 (2016). [CrossRef]  

59. M. Englhard, C. Klemp, M. Behringer, A. Rudolph, O. Skibitzki, P. Zaumseil, and T. Schroeder, “Characterization of reclaimed GaAs substrates and investigation of reuse for thin film InGaAlP LED epitaxial growth,” J. Appl. Phys. 120(4), 045301 (2016). [CrossRef]  

60. K. K. Xu, “Integrated Silicon Directly Modulated Light Source Using p-Well in Standard CMOS Technology,” IEEE Sens. J. 16(16), 6184–6191 (2016). [CrossRef]  

61. J. Li, C. Li, M. Xu, Z. Ji, K. Shi, X. Xu, H. Li, and X. Xu, ““W-shaped” injection current dependence of electroluminescence linewidth in green InGaN/GaN-based LED grown on silicon substrate,” Opt. Express 25(20), A871–A879 (2017). [CrossRef]  

62. R. H. Horng, S. Sinha, C. P. Lee, H. A. Feng, C. Y. Chung, and C. W. Tu, “Composite metal substrate for thin film AlGaInP LED applications,” Opt. Express 27(8), A397–A403 (2019). [CrossRef]  

63. R. H. Horng, H. Y. Chien, K. Y. Chen, W. Y. Tseng, Y. T. Tsai, and F. G. Tarntair, “Development and Fabrication of AlGaInP-Based Flip-Chip Micro-LEDs,” IEEE J. Electron Devices Soc. 6, 475–479 (2018). [CrossRef]  

64. G. Gustafsson, Y. Cao, G. Treacy, F. Klavetter, N. Colaneri, and A. Heeger, “Flexible light-emitting diodes made from soluble conducting polymers,” Nature 357(6378), 477–479 (1992). [CrossRef]  

65. T. H. Kim, K. S. Cho, E. K. Lee, S. J. Lee, J. Chae, J. W. Kim, D. H. Kim, J. Y. Kwon, G. Amaratunga, S. Y. Lee, B. L. Choi, Y. Kuk, J. M. Kim, and K. Kim, “Full-colour quantum dot displays fabricated by transfer printing,” Nat. Photonics 5(3), 176–182 (2011). [CrossRef]  

66. M. A. Meitl, Y. X. Zhou, A. Gaur, S. Jeon, M. L. Usrey, M. S. Strano, and J. A. Rogers, “Solution casting and transfer printing single-walled carbon nanotube films,” Nano Lett. 4(9), 1643–1647 (2004). [CrossRef]  

67. M. J. Allen, V. C. Tung, L. Gomez, Z. Xu, L. M. Chen, K. S. Nelson, C. W. Zhou, R. B. Kaner, and Y. Yang, “Soft Transfer Printing of Chemically Converted Graphene,” Adv. Mater. 21(20), 2098–2102 (2009). [CrossRef]  

68. B. H. Kim, S. Nam, N. Oh, S. Y. Cho, K. J. Yu, C. H. Lee, J. Zhang, K. Deshpande, P. Trefonas, J. H. Kim, J. Lee, J. H. Shin, Y. Yu, J. B. Lim, S. M. Won, Y. K. Cho, N. H. Kim, K. J. Seo, H. Lee, T. I. Kim, M. Shim, and J. A. Rogers, “Multilayer Transfer Printing for Pixelated, Multicolor Quantum Dot Light-Emitting Diodes,” ACS Nano 10(5), 4920–4925 (2016). [CrossRef]  

69. S. Li, B. N. Peele, C. M. Larson, H. C. Zhao, and R. F. Shepherd, “A Stretchable Multicolor Display and Touch Interface Using Photopatterning and Transfer Printing,” Adv. Mater. 28(44), 9770–9775 (2016). [CrossRef]  

70. M. K. Choi, I. Park, D. C. Kim, E. Joh, O. K. Park, J. Kim, M. Kim, C. Choi, J. Yang, K. W. Cho, J. H. Hwang, J. M. Nam, T. Hyeon, J. H. Kim, and D. H. Kim, “Thermally Controlled, Patterned Graphene Transfer Printing for Transparent and Wearable Electronic/Optoelectronic System,” Adv. Funct. Mater. 25(46), 7109–7118 (2015). [CrossRef]  

71. T. S. Pan, M. Pharr, Y. J. Ma, R. Ning, Z. Yan, R. X. Xu, X. Feng, Y. G. Huang, and J. A. Rogers, “Experimental and Theoretical Studies of Serpentine Interconnects on Ultrathin Elastomers for Stretchable Electronics,” Adv. Funct. Mater. 27(37), 1702589 (2017). [CrossRef]  

72. Z. Fan, Y. Zhang, Q. Ma, F. Zhang, H. Fu, K. C. Hwang, and Y. Huang, “A finite deformation model of planar serpentine interconnects for stretchable electronics,” Int. J. Solids Struct. 91, 46–54 (2016). [CrossRef]  

73. M. Nasreldin, R. Delattre, B. Marchiori, M. Ramuz, S. Maria, J. L. D. de la Tocnaye, and T. Djenizian, “Microstructured electrodes supported on serpentine interconnects for stretchable electronics,” APL Mater. 7(3), 031507 (2019). [CrossRef]  

74. L. Xiao, C. Zhu, W. Xiong, Y. Huang, and Z. Yin, “The Conformal Design of an Island-Bridge Structure on a Non-Developable Surface for Stretchable Electronics,” Micromachines 9(8), 392 (2018). [CrossRef]  

75. A. M. V. Mohan, N. Kim, Y. Gu, A. J. Bandodkar, J. M. You, R. Kumar, J. F. Kurniawan, S. Xu, and J. Wang, “Merging of Thin- and Thick-Film Fabrication Technologies: Toward Soft Stretchable “Island-Bridge” Devices,” Adv. Mater. Technol. 2(4), 1600284 (2017). [CrossRef]  

76. R. Li, M. Li, Y. W. Su, J. Z. Song, and X. Q. Ni, “An analytical mechanics model for the island-bridge structure of stretchable electronics,” Soft Matter 9(35), 8476–8482 (2013). [CrossRef]  

77. A. Carlson, A. M. Bowen, Y. Huang, R. G. Nuzzo, and J. A. Rogers, “Transfer printing techniques for materials assembly and micro/nanodevice fabrication,” Adv. Mater. 24(39), 5284–5318 (2012). [CrossRef]  

78. D.-H. Kim, N. Lu, R. Ma, Y.-S. Kim, R.-H. Kim, S. Wang, J. Wu, S. M. Won, H. Tao, and A. Islam, “Epidermal electronics,” Science 333(6044), 838–843 (2011). [CrossRef]  

79. S. Choi, H. Lee, R. Ghaffari, T. Hyeon, and D. H. Kim, “Recent advances in flexible and stretchable bio-electronic devices integrated with nanomaterials,” Adv. Mater. 28(22), 4203–4218 (2016). [CrossRef]  

80. H. C. Ko, M. P. Stoykovich, J. Song, V. Malyarchuk, W. M. Choi, C.-J. Yu, J. B. Geddes Iii, J. Xiao, S. Wang, and Y. Huang, “A hemispherical electronic eye camera based on compressible silicon optoelectronics,” Nature 454(7205), 748–753 (2008). [CrossRef]  

81. J. Kim, M. Lee, H. J. Shim, R. Ghaffari, H. R. Cho, D. Son, Y. H. Jung, M. Soh, C. Choi, and S. Jung, “Stretchable silicon nanoribbon electronics for skin prosthesis,” Nat. Commun. 5(1), 5747 (2014). [CrossRef]  

82. Y. Liu, M. Pharr, and G. A. Salvatore, “Lab-on-skin: a review of flexible and stretchable electronics for wearable health monitoring,” ACS Nano 11(10), 9614–9635 (2017). [CrossRef]  

83. M. Kaltenbrunner, G. Kettlgruber, C. Siket, R. Schwödiauer, and S. Bauer, “Arrays of ultracompliant electrochemical dry gel cells for stretchable electronics,” Adv. Mater. 22(18), 2065–2067 (2010). [CrossRef]  

84. W. Chee, H. Lim, Z. Zainal, N. Huang, I. Harrison, and Y. Andou, “Flexible graphene-based supercapacitors: a review,” J. Phys. Chem. C 120(8), 4153–4172 (2016). [CrossRef]  

85. M. Kaltenbrunner, M. S. White, E. D. Głowacki, T. Sekitani, T. Someya, N. S. Sariciftci, and S. Bauer, “Ultrathin and lightweight organic solar cells with high flexibility,” Nat. Commun. 3(1), 770 (2012). [CrossRef]  

86. J. Wang, S. Li, F. Yi, Y. Zi, J. Lin, X. Wang, Y. Xu, and Z. L. Wang, “Sustainably powering wearable electronics solely by biomechanical energy,” Nat. Commun. 7(1), 12744 (2016). [CrossRef]  

87. M. F. El-Kady, V. Strong, S. Dubin, and R. B. Kaner, “Laser scribing of high-performance and flexible graphene-based electrochemical capacitors,” Science 335(6074), 1326–1330 (2012). [CrossRef]  

88. M. K. Choi, J. Yang, T. Hyeon, and D.-H. Kim, “Flexible quantum dot light-emitting diodes for next-generation displays,” npj Flex Electron 2(1), 10 (2018). [CrossRef]  

89. J. Kim, A. S. Campbell, B. E.-F. de Ávila, and J. Wang, “Wearable biosensors for healthcare monitoring,” Nat. Biotechnol. 37(4), 389 (2019). [CrossRef]  

90. J. R. Sempionatto, T. Nakagawa, A. Pavinatto, S. T. Mensah, S. Imani, P. Mercier, and J. Wang, “Eyeglasses based wireless electrolyte and metabolite sensor platform,” Lab Chip 17(10), 1834–1842 (2017). [CrossRef]  

91. A. J. Bandodkar, W. Jia, C. Yardımcı, X. Wang, J. Ramirez, and J. Wang, “Tattoo-based noninvasive glucose monitoring: a proof-of-concept study,” Anal. Chem. 87(1), 394–398 (2015). [CrossRef]  

92. M. S. Mannoor, H. Tao, J. D. Clayton, A. Sengupta, D. L. Kaplan, R. R. Naik, N. Verma, F. G. Omenetto, and M. C. McAlpine, “Graphene-based wireless bacteria detection on tooth enamel,” Nat. Commun. 3(1), 763 (2012). [CrossRef]  

93. A. Koh, D. Kang, Y. Xue, S. Lee, R. M. Pielak, J. Kim, T. Hwang, S. Min, A. Banks, and P. Bastien, “A soft, wearable microfluidic device for the capture, storage, and colorimetric sensing of sweat,” Sci. Transl. Med. 8(366), 366ra165 (2016). [CrossRef]  

94. H. Lee, T. K. Choi, Y. B. Lee, H. R. Cho, R. Ghaffari, L. Wang, H. J. Choi, T. D. Chung, N. Lu, and T. Hyeon, “A graphene-based electrochemical device with thermoresponsive microneedles for diabetes monitoring and therapy,” Nat. Nanotechnol. 11(6), 566–572 (2016). [CrossRef]  

95. S. Emaminejad, W. Gao, E. Wu, Z. A. Davies, H. Y. Y. Nyein, S. Challa, S. P. Ryan, H. M. Fahad, K. Chen, and Z. Shahpar, “Autonomous sweat extraction and analysis applied to cystic fibrosis and glucose monitoring using a fully integrated wearable platform,” Proc. Natl. Acad. Sci. 114(18), 4625–4630 (2017). [CrossRef]  

96. I. Jeerapan, J. R. Sempionatto, A. Pavinatto, J.-M. You, and J. Wang, “Stretchable biofuel cells as wearable textile-based self-powered sensors,” J. Mater. Chem. A 4(47), 18342–18353 (2016). [CrossRef]  

97. R. D. Munje, S. Muthukumar, B. Jagannath, and S. Prasad, “A new paradigm in sweat based wearable diagnostics biosensors using Room Temperature Ionic Liquids (RTILs),” Sci. Rep. 7(1), 1950 (2017). [CrossRef]  

98. W. Gao, S. Emaminejad, H. Y. Y. Nyein, S. Challa, K. Chen, A. Peck, H. M. Fahad, H. Ota, H. Shiraki, and D. Kiriya, “Fully integrated wearable sensor arrays for multiplexed in situ perspiration analysis,” Nature 529(7587), 509–514 (2016). [CrossRef]  

99. S. Imani, A. J. Bandodkar, A. V. Mohan, R. Kumar, S. Yu, J. Wang, and P. P. Mercier, “A wearable chemical–electrophysiological hybrid biosensing system for real-time health and fitness monitoring,” Nat. Commun. 7(1), 11650 (2016). [CrossRef]  

100. M. Senior, Novartis signs up for Google smart lens (Nature Publishing Group, 2014).

101. Z. Bao and X. Chen, “Flexible and Stretchable Devices,” Adv. Mater. 28(22), 4177–4179 (2016). [CrossRef]  

102. G. Park, H. J. Chung, K. Kim, S. A. Lim, J. Kim, Y. S. Kim, Y. Liu, W. H. Yeo, R. H. Kim, S. S. Kim, J. S. Kim, Y. H. Jung, T. I. Kim, C. Yee, J. A. Rogers, and K. M. Lee, “Immunologic and tissue biocompatibility of flexible/stretchable electronics and optoelectronics,” Adv. Healthcare Mater. 3(4), 515–525 (2014). [CrossRef]  

103. S Mack, M. A. Meitl, A. J. Baca, Z. T. Zhu, and J. A. Rogers, “Mechanically flexible thin-film transistors that use ultrathin ribbons of silicon derived from bulk wafers,” Appl. Phys. Lett. 88, 213101 (2006). [CrossRef]  

104. J. J. Hu, L. Li, H. T. Lin, P. Zhang, W. D. Zhou, and Z. Q. Ma, “Flexible integrated photonics: where materials, mechanics and optics meet [Invited],” Opt. Mater. Express 3(9), 1313–1331 (2013). [CrossRef]  

105. J. Song, “Mechanics of stretchable electronics,” Curr. Opin. Solid State Mater. Sci. 19(3), 160–170 (2015). [CrossRef]  

106. Y. Sun, W. M. Choi, H. Jiang, Y. Y. Huang, and J. A. Rogers, “Controlled buckling of semiconductor nanoribbons for stretchable electronics,” Nat. Nanotechnol. 1(3), 201–207 (2006). [CrossRef]  

107. A. J. Baca, M. A. Meitl, H. C. Ko, S. Mack, H. S. Kim, J. Dong, P. M. Ferreira, and J. A. Rogers, “Printable Single-Crystal Silicon Micro/Nanoscale Ribbons, Platelets and Bars Generated from Bulk Wafers,” Adv. Funct. Mater. 17(16), 3051–3062 (2007). [CrossRef]  

108. H.-C. Seo, I. Petrov, H. Jeong, P. Chapman, and K. Kim, “Elastic buckling of AlN ribbons on elastomeric substrate,” Appl. Phys. Lett. 94(9), 092104 (2009). [CrossRef]  

109. X. Feng, B. D. Yang, Y. Liu, Y. Wang, C. Dagdeviren, Z. Liu, A. Carlson, J. Li, Y. Huang, and J. A. Rogers, “Stretchable ferroelectric nanoribbons with wavy configurations on elastomeric substrates,” ACS Nano 5(4), 3326–3332 (2011). [CrossRef]  

110. J. A. Fan, W. H. Yeo, Y. Su, Y. Hattori, W. Lee, S. Y. Jung, Y. Zhang, Z. Liu, H. Cheng, L. Falgout, M. Bajema, T. Coleman, D. Gregoire, R. J. Larsen, Y. Huang, and J. A. Rogers, “Fractal design concepts for stretchable electronics,” Nat. Commun. 5(1), 3266 (2014). [CrossRef]  

111. C.-H. Hsueh, “Modeling of elastic deformation of multilayers due to residual stresses and external bending,” J. Appl. Phys. 91(12), 9652 (2002). [CrossRef]  

112. D. Y. Khang, H. Jiang, Y. Huang, and J. A. Rogers, “A stretchable form of single-crystal silicon for high-performance electronics on rubber substrates,” Science 311(5758), 208–212 (2006). [CrossRef]  

113. X. Fu, C. Su, Q. Fu, X. Zhu, R. Zhu, C. Liu, Z. Liao, J. Xu, W. Guo, J. Feng, J. Li, and D. Yu, “Tailoring exciton dynamics by elastic strain-gradient in semiconductors,” Adv. Mater. 26(16), 2572–2579 (2014). [CrossRef]  

114. Z. Liu, J. Xu, D. Chen, and G. Shen, “Flexible electronics based on inorganic nanowires,” Chem. Soc. Rev. 44(1), 161–192 (2015). [CrossRef]  

115. H. Liu, H. Tang, M. Fang, W. Si, Q. Zhang, Z. Huang, L. Gu, W. Pan, J. Yao, C. Nan, and H. Wu, “2D Metals by Repeated Size Reduction,” Adv. Mater. 28(37), 8170–8176 (2016). [CrossRef]  

116. R. L. Petritz, “Theory of Photoconductivity in Semiconductor Films,” Phys. Rev. 104(6), 1508–1516 (1956). [CrossRef]  

117. G. Konstantatos and E. H. Sargent, “Solution-processed quantum dot photodetectors,” Proc. IEEE 97(10), 1666–1683 (2009). [CrossRef]  

118. Y. Zhao and K. Zhu, “Organic–inorganic hybrid lead halide perovskites for optoelectronic and electronic applications,” Chem. Soc. Rev. 45(3), 655–689 (2016). [CrossRef]  

119. H. Chen, H. Liu, Z. Zhang, K. Hu, and X. Fang, “Nanostructured photodetectors: from ultraviolet to terahertz,” Adv. Mater. 28(3), 403–433 (2016). [CrossRef]  

120. K. J. Baeg, M. Binda, D. Natali, M. Caironi, and Y. Y. Noh, “Organic light detectors: photodiodes and phototransistors,” Adv. Mater. 25(31), 4267–4295 (2013). [CrossRef]  

121. K. Deng and L. Li, “CdS nanoscale photodetectors,” Adv. Mater. 26(17), 2619–2635 (2014). [CrossRef]  

122. J. Jie, W. Zhang, I. Bello, C.-S. Lee, and S.-T. Lee, “One-dimensional II–VI nanostructures: synthesis, properties and optoelectronic applications,” Nano today 5(4), 313–336 (2010). [CrossRef]  

123. Y. B. Wang, L. F. Wang, H. J. Joyce, Q. Gao, X. Z. Liao, Y. W. Mai, H. H. Tan, J. Zou, S. P. Ringer, H. J. Gao, and C. Jagadish, “Super deformability and Young's modulus of GaAs nanowires,” Adv. Mater. 23(11), 1356–1360 (2011). [CrossRef]  

124. S. Yang, C. Wang, H. Sahin, H. Chen, Y. Li, S.-S. Li, A. Suslu, F. M. Peeters, Q. Liu, and J. Li, “Tuning the optical, magnetic, and electrical properties of ReSe2 by nanoscale strain engineering,” Nano Lett. 15(3), 1660–1666 (2015). [CrossRef]  

125. G. Signorello, S. Karg, M.T. Björk, B. Gotsmann, and H. Riel, “Tuning the light emission from GaAs nanowires over 290 meV with uniaxial strain,” Nano Lett. 13(3), 917–924 (2013). [CrossRef]  

126. J. Qi, X. Qian, L. Qi, J. Feng, D. Shi, and J. Li, “Strain-engineering of band gaps in piezoelectric boron nitride nanoribbons,” Nano Lett. 12(3), 1224–1228 (2012). [CrossRef]  

127. S. B. Desai, G. Seol, J. S. Kang, H. Fang, C. Battaglia, R. Kapadia, J. W. Ager, J. Guo, and A. Javey, “Strain-induced indirect to direct bandgap transition in multilayer WSe2,” Nano Lett. 14(8), 4592–4597 (2014). [CrossRef]  

128. H. J. Conley, B. Wang, J. I. Ziegler, R. F. Haglund Jr, S. T. Pantelides, and K. I. Bolotin, “Bandgap engineering of strained monolayer and bilayer MoS2,” Nano Lett. 13(8), 3626–3630 (2013). [CrossRef]  

129. S. I. Park, A. P. Le, J. Wu, Y. Huang, X. Li, and J. A. Rogers, “Light Emission Characteristics and Mechanics of Foldable Inorganic Light-Emitting Diodes,” Adv. Mater. 22(28), 3062–3066 (2010). [CrossRef]  

130. E. T. Roche, M. A. Horvath, I. Wamala, A. Alazmani, S.-E. Song, W. Whyte, Z. Machaidze, C. J. Payne, J. C. Weaver, and G. Fishbein, “Soft robotic sleeve supports heart function,” Sci. Transl. Med. 9(373), eaaf3925 (2017). [CrossRef]  

131. G. A. Salvatore, J. Sülzle, F. Dalla Valle, G. Cantarella, F. Robotti, P. Jokic, S. Knobelspies, A. Daus, L. Büthe, and L. Petti, “Biodegradable and highly deformable temperature sensors for the internet of things,” Adv. Funct. Mater. 27(35), 1702390 (2017). [CrossRef]  

132. I. Johnston, D. McCluskey, C. Tan, and M. Tracey, “Mechanical characterization of bulk Sylgard 184 for microfluidics and microengineering,” J. Micromech. Microeng. 24(3), 035017 (2014). [CrossRef]  

133. S.-J. Kim, D.-S. Lee, I.-G. Kim, D.-W. Sohn, J.-Y. Park, B.-K. Choi, and S.-W. Kim, “Evaluation of the biocompatibility of a coating material for an implantable bladder volume sensor,” The Kaohsiung journal of medical sciences 28(3), 123–129 (2012). [CrossRef]  

134. J. del Valle, N. de la Oliva, M. Müller, T. Stieglitz, and X. Navarro, “Biocompatibility evaluation of parylene C and polyimide as substrates for peripheral nerve interfaces,” in 2015 7th International IEEE/EMBS Conference on Neural Engineering (NER), (IEEE, 2015), 442–445.

135. Y. Chen, B. Lu, Y. Chen, and X. Feng, “Breathable and stretchable temperature sensors inspired by skin,” Sci. Rep. 5(1), 11505 (2015). [CrossRef]  

136. Y. Chen, S. Lu, S. Zhang, Y. Li, Z. Qu, Y. Chen, B. Lu, X. Wang, and X. Feng, “Skin-like biosensor system via electrochemical channels for noninvasive blood glucose monitoring,” Sci. Adv. 3(12), e1701629 (2017). [CrossRef]  

137. C. Mills, J. Escarré, E. Engel, E. Martinez, A. Errachid, J. Bertomeu, J. Andreu, J. A. Planell, and J. Samitier, “Micro-and nanostructuring of poly (ethylene-2, 6-naphthalate) surfaces, for biomedical applications, using polymer replication techniques,” Nanotechnology 16(4), 369–375 (2005). [CrossRef]  

138. H. Tao, S.-W. Hwang, B. Marelli, B. An, J. E. Moreau, M. Yang, M. A. Brenckle, S. Kim, D. L. Kaplan, and J. A. Rogers, “Silk-based resorbable electronic devices for remotely controlled therapy and in vivo infection abatement,” Proc. Natl. Acad. Sci. 111(49), 17385–17389 (2014). [CrossRef]  

139. P. Gentile, V. Chiono, I. Carmagnola, and P. Hatton, “An overview of poly (lactic-co-glycolic) acid (PLGA)-based biomaterials for bone tissue engineering,” Int. J. Mol. Sci. 15(3), 3640–3659 (2014). [CrossRef]  

140. A. Ignatius and L. E. Claes, “In vitro biocompatibility of bioresorbable polymers: poly (L, DL-lactide) and poly (L-lactide-co-glycolide),” Biomaterials 17(8), 831–839 (1996). [CrossRef]  

141. I. Tubia, M. Mujika, J. Artieda, M. Valencia, and S. Arana, “Soft polymer sensor for recording surface cortical activity in freely moving rodents,” Sens. Actuators, A 251, 241–247 (2016). [CrossRef]  

142. S. Lu, X. Wang, Q. Lu, X. Zhang, J. A. Kluge, N. Uppal, F. Omenetto, and D. L. Kaplan, “Insoluble and flexible silk films containing glycerol,” Biomacromolecules 11(1), 143–150 (2010). [CrossRef]  

143. S. Kanazawa, Y. Kusaka, N. Yamamoto, and H. Ushijima, “Improved Transfer Process for the Fully Additive Manufacturing of a Conductive Layer-Stacked Polymeric Cantilever,” Mater. Sci. Appl. 10(01), 45–52 (2019). [CrossRef]  

144. Y. Sun and J. A. Rogers, “Inorganic Semiconductors for Flexible Electronics,” Adv. Mater. 19(15), 1897–1916 (2007). [CrossRef]  

145. J. Justice, C. Bower, M. Meitl, M. B. Mooney, M. A. Gubbins, and B. Corbett, “Wafer-scale integration of group III–V lasers on silicon using transfer printing of epitaxial layers,” Nat. Photonics 6(9), 610–614 (2012). [CrossRef]  

146. Y. Sato, M. Sakaki, M. Katayama, M. Higuma, M. Kudo, and K. Moriya, “Image-transfer medium for ink-jet printing, transfer printing process using the same, and transfer printing cloth,” (Google Patents, 2002).

147. T.-H. Kim, A. Carlson, J.-H. Ahn, S. M. Won, S. Wang, Y. Huang, and J. A. Rogers, “Kinetically controlled, adhesiveless transfer printing using microstructured stamps,” Appl. Phys. Lett. 94(11), 113502 (2009). [CrossRef]  

148. M. A. Meitl, Y. Zhou, A. Gaur, S. Jeon, M. L. Usrey, M. S. Strano, and J. A. Rogers, “Solution casting and transfer printing single-walled carbon nanotube films,” Nano Lett. 4(9), 1643–1647 (2004). [CrossRef]  

149. L. Chen, P. Degenaar, and D. D. Bradley, “Polymer transfer printing: application to layer coating, pattern definition, and diode dark current blocking,” Adv. Mater. 20(9), 1679–1683 (2008). [CrossRef]  

150. K. H. Yim, Z. Zheng, Z. Liang, R. H. Friend, W. T. Huck, and J. S. Kim, “Efficient Conjugated-Polymer Optoelectronic Devices Fabricated by Thin-Film Transfer-Printing Technique,” Adv. Funct. Mater. 18(7), 1012–1019 (2008). [CrossRef]  

151. M. A. Meitl, Z. T. Zhu, V. Kumar, K. J. Lee, X. Feng, Y. Y. Huang, I. Adesida, R. G. Nuzzo, and J. A. Rogers, “Transfer printing by kinetic control of adhesion to an elastomeric stamp,” Nat. Mater. 5(1), 33–38 (2006). [CrossRef]  

152. A. Carlson, A. M. Bowen, Y. Huang, R. G. Nuzzo, and J. A. Rogers, “Transfer printing techniques for materials assembly and micro/nanodevice fabrication,” Adv. Mater. 24(39), 5284–5318 (2012). [CrossRef]  

153. J. J. Wang, J. X. Xie, C. Y. Zong, X. Han, H. P. Ji, J. X. Zhao, and C. H. Lu, “Surface treatment-assisted switchable transfer printing on polydimethylsiloxane films,” J. Mater. Chem. C 4(16), 3467–3476 (2016). [CrossRef]  

154. H. Lee, D. S. Um, Y. Lee, S. Lim, H. J. Kim, and H. Ko, “Octopus-Inspired Smart Adhesive Pads for Transfer Printing of Semiconducting Nanomembranes,” Adv. Mater. 28(34), 7457–7465 (2016). [CrossRef]  

155. Y. Huang, N. Zheng, Z. Cheng, Y. Chen, B. Lu, T. Xie, and X. Feng, “Direct Laser Writing-Based Programmable Transfer Printing via Bioinspired Shape Memory Reversible Adhesive,” ACS Appl. Mater. Interfaces 8(51), 35628–35633 (2016). [CrossRef]  

156. S. I. Park, Y. J. Xiong, R. H. Kim, P. Elvikis, M. Meitl, D. H. Kim, J. Wu, J. Yoon, C. J. Yu, Z. J. Liu, Y. G. Huang, K. Hwang, P. Ferreira, X. L. Li, K. Choquette, and J. A. Rogers, “Printed Assemblies of Inorganic Light-Emitting Diodes for Deformable and Semitransparent Displays,” Science 325(5943), 977–981 (2009). [CrossRef]  

157. H. S. Kim, E. Brueckner, J. Z. Song, Y. H. Li, S. Kim, C. F. Lu, J. Sulkin, K. Choquette, Y. G. Huang, R. G. Nuzzo, and J. A. Rogers, “Unusual strategies for using indium gallium nitride grown on silicon (111) for solid-state lighting,” Proc. Natl. Acad. Sci. U. S. A. 108(25), 10072–10077 (2011). [CrossRef]  

158. H. E. Lee, D. Lee, T. I. Lee, J. H. Shin, G. M. Choi, C. Kim, S. H. Lee, J. H. Lee, Y. H. Kim, S. M. Kang, S. H. Park, I. S. Kang, T. S. Kim, B. S. Bae, and K. J. Lee, “Wireless powered wearable micro light-emitting diodes,” Nano Energy 55, 454–462 (2019). [CrossRef]  

159. C. A. Bower, M. A. Meitl, B. Raymond, E. Radauscher, R. Cok, S. Bonafede, D. Gomez, T. Moore, C. Prevatte, B. Fisher, R. Rotzoll, G. A. Melnik, A. Fecioru, and A. J. Trindade, “Emissive displays with transfer-printed assemblies of 8 µm × 15 µm inorganic light-emitting diodes,” Photonics Res. 5(2), A23 (2017). [CrossRef]  

160. A. Nadarajah, R. C. Word, J. Meiss, and R. Könenkamp, “Flexible inorganic nanowire light-emitting diode,” Nano Lett. 8(2), 534–537 (2008). [CrossRef]  

161. M. Koo, S. Y. Park, and K. J. Lee, “Biointegrated flexible inorganic light emitting diodes,” NDD 2012(1), 5–15 (2012). [CrossRef]  

162. Y. He, J.-a. Wang, W. Zhang, J. Song, C. Pei, and X. Chen, “Zno-nanowires/pani inorganic/organic heterostructure light-emitting diode,” J. Nanosci. Nanotechnol. 10(11), 7254–7257 (2010). [CrossRef]  

163. M. Vosgueritchian, J. B.-H. Tok, and Z. Bao, “Stretchable LEDs: Light-emitting electronic skin,” Nat. Photonics 7(10), 769–771 (2013). [CrossRef]  

164. J.-A. Jeong, H.-S. Shin, K.-H. Choi, and H.-K. Kim, “Flexible Al-doped ZnO films grown on PET substrates using linear facing target sputtering for flexible OLEDs,” J. Phys. D: Appl. Phys. 43(46), 465403 (2010). [CrossRef]  

165. S.-M. Lee, J. H. Kwon, S. Kwon, and K. C. Choi, “A review of flexible OLEDs toward highly durable unusual displays,” IEEE Trans. Electron Devices 64(5), 1922–1931 (2017). [CrossRef]  

166. T.-H. Han, Y. Lee, M.-R. Choi, S.-H. Woo, S.-H. Bae, B. H. Hong, J.-H. Ahn, and T.-W. Lee, “Extremely efficient flexible organic light-emitting diodes with modified graphene anode,” Nat. Photonics 6(2), 105–110 (2012). [CrossRef]  

167. J.-J. Kim, M.-K. Han, and Y.-Y. Noh, “Flexible OLEDs and organic electronics,” Semicond. Sci. Technol. 26(3), 030301 (2011). [CrossRef]  

168. F. L. Jakobsson, X. Crispin, L. Lindell, A. Kanciurzewska, M. Fahlman, W. R. Salaneck, and M. Berggren, “Towards all-plastic flexible light emitting diodes,” Chem. Phys. Lett. 433(1-3), 110–114 (2006). [CrossRef]  

169. M. S. White, M. Kaltenbrunner, E. D. Głowacki, K. Gutnichenko, G. Kettlgruber, I. Graz, S. Aazou, C. Ulbricht, D. A. Egbe, and M. C. Miron, “Ultrathin, highly flexible and stretchable PLEDs,” Nat. Photonics 7(10), 811–816 (2013). [CrossRef]  

170. N. Tessler, V. Medvedev, M. Kazes, S. Kan, and U. Banin, “Efficient near-infrared polymer nanocrystal light-emitting diodes,” Science 295(5559), 1506–1508 (2002). [CrossRef]  

171. X. Yang, E. Mutlugun, C. Dang, K. Dev, Y. Gao, S. T. Tan, X. W. Sun, and H. V. Demir, “Highly flexible, electrically driven, top-emitting, quantum dot light-emitting stickers,” ACS nano 8(8), 8224–8231 (2014). [CrossRef]  

172. R. Liang, D. Yan, R. Tian, X. Yu, W. Shi, C. Li, M. Wei, D. G. Evans, and X. Duan, “Quantum dots-based flexible films and their application as the phosphor in white light-emitting diodes,” Chem. Mater. 26(8), 2595–2600 (2014). [CrossRef]  

173. M. Choi, B. Jang, W. Lee, S. Lee, T. W. Kim, H. J. Lee, J. H. Kim, and J. H. Ahn, “Stretchable Active Matrix Inorganic Light-Emitting Diode Display Enabled by Overlay-Aligned Roll-Transfer Printing,” Adv. Funct. Mater. 27(11), 1606005 (2017). [CrossRef]  

174. Y. Cao, G. G. Zhang, Y. C. Zhang, M. K. Yue, Y. Chen, S. S. Cai, T. Xie, and X. Feng, “Direct Fabrication of Stretchable Electronics on a Polymer Substrate with Process-Integrated Programmable Rigidity,” Adv. Funct. Mater. 28(50), 1804604 (2018). [CrossRef]  

175. R. H. Kim, M. H. Bae, D. G. Kim, H. Cheng, B. H. Kim, D. H. Kim, M. Li, J. Wu, F. Du, H. S. Kim, S. Kim, D. Estrada, S. W. Hong, Y. Huang, E. Pop, and J. A. Rogers, “Stretchable, transparent graphene interconnects for arrays of microscale inorganic light emitting diodes on rubber substrates,” Nano Lett. 11(9), 3881–3886 (2011). [CrossRef]  

176. J. H. Seo, K. Zhang, M. Kim, D. Zhao, H. Yang, W. Zhou, and Z. Ma, “Flexible Phototransistors Based on Single-Crystalline Silicon Nanomembranes,” Adv. Opt. Mater. 4(1), 120–125 (2016). [CrossRef]  

177. J. Lee, J. Wu, M. Shi, J. Yoon, S. I. Park, M. Li, Z. Liu, Y. Huang, and J. A. Rogers, “Stretchable GaAs photovoltaics with designs that enable high areal coverage,” Adv. Mater. 23(8), 986–991 (2011). [CrossRef]  

178. J. Wu and L. Y. Lin, “A flexible nanocrystal photovoltaic ultraviolet photodetector on a plant membrane,” Adv. Opt. Mater. 3(11), 1530–1536 (2015). [CrossRef]  

179. C. Soci, A. Zhang, B. Xiang, S. A. Dayeh, D. Aplin, J. Park, X. Bao, Y.-H. Lo, and D. Wang, “ZnO nanowire UV photodetectors with high internal gain,” Nano Lett. 7(4), 1003–1009 (2007). [CrossRef]  

180. L. Hu, J. Yan, M. Liao, L. Wu, and X. Fang, “Ultrahigh external quantum efficiency from thin SnO2 nanowire ultraviolet photodetectors,” Small 7(8), 1012–1017 (2011). [CrossRef]  

181. W. Tian, T. Zhai, C. Zhang, S. L. Li, X. Wang, F. Liu, D. Liu, X. Cai, K. Tsukagoshi, and D. Golberg, “Low-Cost Fully Transparent Ultraviolet Photodetectors Based on Electrospun ZnO-SnO2 Heterojunction Nanofibers,” Adv. Mater. 25(33), 4625–4630 (2013). [CrossRef]  

182. J.-M. Wu and C.-H. Kuo, “Ultraviolet photodetectors made from SnO2 nanowires,” Thin solid films 517(14), 3870–3873 (2009). [CrossRef]  

183. Z. Lou, L. Li, and G. Shen, “High-performance rigid and flexible ultraviolet photodetectors with single-crystalline ZnGa 2 O 4 nanowires,” Nano Res. 8(7), 2162–2169 (2015). [CrossRef]  

184. G. Shen and D. Chen, “One-dimensional nanostructures for photodetectors,” Recent Pat. Nanotechnol. 4(1), 20–31 (2010). [CrossRef]  

185. C.-L. Hsu, Y.-R. Lin, S.-J. Chang, T.-S. Lin, S.-Y. Tsai, and I.-C. Chen, “Vertical ZnO/ZnGa2O4 core–shell nanorods grown on ZnO/glass templates by reactive evaporation,” Chem. Phys. Lett. 411(1-3), 221–224 (2005). [CrossRef]  

186. T. Zhai, X. Fang, M. Liao, X. Xu, H. Zeng, B. Yoshio, and D. Golberg, “A comprehensive review of one-dimensional metal-oxide nanostructure photodetectors,” Sensors 9(8), 6504–6529 (2009). [CrossRef]  

187. Z. Zheng, L. Gan, H. Li, Y. Ma, Y. Bando, D. Golberg, and T. Zhai, “A Fully Transparent and Flexible Ultraviolet–Visible Photodetector Based on Controlled Electrospun ZnO-CdO Heterojunction Nanofiber Arrays,” Adv. Funct. Mater. 25(37), 5885–5894 (2015). [CrossRef]  

188. B. Yin, H. Zhang, Y. Qiu, Y. Luo, Y. Zhao, and L. Hu, “Piezo-phototronic effect enhanced self-powered and broadband photodetectors based on Si/ZnO/CdO three-component heterojunctions,” Nano energy 40, 440–446 (2017). [CrossRef]  

189. Z. Zheng, L. Gan, H. Li, Y. Ma, Y. Bando, D. Golberg, and T. Zhai, “Photodetectors: A Fully Transparent and Flexible Ultraviolet–Visible Photodetector Based on Controlled Electrospun ZnO-CdO Heterojunction Nanofiber Arrays (Adv. Funct. Mater. 37/2015),” Adv. Funct. Mater. 25(37), 5877 (2015). [CrossRef]  

190. F. Zhang, S. Niu, W. Guo, G. Zhu, Y. Liu, X. Zhang, and Z. L. Wang, “Piezo-phototronic effect enhanced visible/UV photodetector of a carbon-fiber/ZnO-CdS double-shell microwire,” Acs Nano 7(5), 4537–4544 (2013). [CrossRef]  

191. C. Yan, J. Wang, X. Wang, W. Kang, M. Cui, C. Y. Foo, and P. S. Lee, “An intrinsically stretchable nanowire photodetector with a fully embedded structure,” Adv. Mater. 26(6), 943–950 (2014). [CrossRef]  

192. S. S. Yoon and D. Y. Khang, “Stretchable, Bifacial Si-Organic Hybrid Solar Cells by Vertical Array of Si Micropillars Embedded into Elastomeric Substrates,” ACS Appl. Mater. Interfaces 11(3), 3290–3298 (2019). [CrossRef]  

193. M. Park, H. J. Kim, I. Jeong, J. Lee, H. Lee, H. J. Son, D.-E. Kim, and M. J. Ko, “Mechanically Recoverable and Highly Efficient Perovskite Solar Cells: Investigation of Intrinsic Flexibility of Organic-Inorganic Perovskite,” Adv. Energy Mater. 5(22), 1501406 (2015). [CrossRef]  

194. X. Wang, W. Song, B. Liu, G. Chen, D. Chen, C. Zhou, and G. Shen, “High-performance organic-inorganic hybrid photodetectors based on P3HT: CdSe nanowire heterojunctions on rigid and flexible substrates,” Adv. Funct. Mater. 23(9), 1202–1209 (2013). [CrossRef]  

195. L. Li, F. Zhang, J. Wang, Q. An, Q. Sun, W. Wang, J. Zhang, and F. Teng, “Achieving EQE of 16,700% in P3HT: PC 71 BM based photodetectors by trap-assisted photomultiplication,” Sci. Rep. 5(1), 9181 (2015). [CrossRef]  

196. H. Wei, Y. Fang, Y. Yuan, L. Shen, and J. Huang, “Trap engineering of CdTe nanoparticle for high gain, fast response, and low noise P3HT: CdTe nanocomposite photodetectors,” Adv. Mater. 27(34), 4975–4981 (2015). [CrossRef]  

197. H. Li, T. Chen, S. Hu, X. Li, Y. Liu, P. Lee, X. Wang, H. Li, and G. Lo, “Highly spectrum-selective ultraviolet photodetector based on p-NiO/n-IGZO thin film heterojunction structure,” Opt. Express 23(21), 27683–27689 (2015). [CrossRef]  

198. Z. Pei, H.-C. Lai, J.-Y. Wang, W.-H. Chiang, and C.-H. Chen, “High-responsivity and high-sensitivity graphene dots/a-IGZO thin-film phototransistor,” IEEE Electron Device Lett. 36(1), 44–46 (2015). [CrossRef]  

199. J. Yu, K. Javaid, L. Liang, W. Wu, Y. Liang, A. Song, H. Zhang, W. Shi, T.-C. Chang, and H. Cao, “High-Performance Visible-Blind Ultraviolet Photodetector Based on IGZO TFT Coupled with p–n Heterojunction,” ACS Appl. Mater. Interfaces 10(9), 8102–8109 (2018). [CrossRef]  

200. Y. Jiang, W. J. Zhang, J. S. Jie, X. M. Meng, X. Fan, and S. T. Lee, “Photoresponse properties of CdSe single-nanoribbon photodetectors,” Adv. Funct. Mater. 17(11), 1795–1800 (2007). [CrossRef]  

201. D. C. Oertel, M. G. Bawendi, A. C. Arango, and V. Bulović, “Photodetectors based on treated CdSe quantum-dot films,” Appl. Phys. Lett. 87(21), 213505 (2005). [CrossRef]  

202. E. Bosman, G. Van Steenberge, B. Van Hoe, J. Missinne, J. Vanfleteren, and P. Van Daele, “Highly Reliable Flexible Active Optical Links,” IEEE Photonics Technol. Lett. 22(5), 287–289 (2010). [CrossRef]  

203. L. Fan, L. T. Varghese, Y. Xuan, J. Wang, B. Niu, and M. Qi, “Direct fabrication of silicon photonic devices on a flexible platform and its application for strain sensing,” Opt. Express 20(18), 20564–20575 (2012). [CrossRef]  

204. L. Li, H. Lin, S. Qiao, Y. Z. Huang, J. Y. Li, J. Michon, T. Gu, C. Alosno-Ramos, L. Vivien, A. Yadav, K. Richardson, N. Lu, and J. Hu, “Monolithically integrated stretchable photonics,” Light: Sci. Appl. 7(2), 17138 (2018). [CrossRef]  

205. H. Cho, S. Han, J. Kwon, J. Jung, H. J. Kim, H. Kim, H. Eom, S. Hong, and S. H. Ko, “Self-assembled stretchable photonic crystal for a tunable color filter,” Opt. Lett. 43(15), 3501–3504 (2018). [CrossRef]  

206. Y. M. Song, Y. Xie, V. Malyarchuk, J. Xiao, I. Jung, K. J. Choi, Z. Liu, H. Park, C. Lu, R. H. Kim, R. Li, K. B. Crozier, Y. Huang, and J. A. Rogers, “Digital cameras with designs inspired by the arthropod eye,” Nature 497(7447), 95–99 (2013). [CrossRef]  

207. K. Zhang, Y. H. Jung, S. Mikael, J. H. Seo, M. Kim, H. Mi, H. Zhou, Z. Xia, W. Zhou, S. Gong, and Z. Ma, “Origami silicon optoelectronics for hemispherical electronic eye systems,” Nat. Commun. 8(1), 1782 (2017). [CrossRef]  

208. C. Liu, Q. Zhang, D. Wang, G. Zhao, X. Cai, L. Li, H. Ding, K. Zhang, H. Wang, D. Kong, L. Yin, L. Liu, G. Zou, L. Zhao, and X. Sheng, “High Performance, Biocompatible Dielectric Thin-Film Optical Filters Integrated with Flexible Substrates and Microscale Optoelectronic Devices,” Adv. Opt. Mater. 6(15), 1800146 (2018). [CrossRef]  

209. T. I. Kim, J. G. McCall, Y. H. Jung, X. Huang, E. R. Siuda, Y. Li, J. Song, Y. M. Song, H. A. Pao, R. H. Kim, C. Lu, S. D. Lee, I. S. Song, G. Shin, R. Al-Hasani, S. Kim, M. P. Tan, Y. Huang, F. G. Omenetto, J. A. Rogers, and M. R. Bruchas, “Injectable, cellular-scale optoelectronics with applications for wireless optogenetics,” Science 340(6129), 211–216 (2013). [CrossRef]  

210. J. Kim, P. Gutruf, A. M. Chiarelli, S. Y. Heo, K. Cho, Z. Xie, A. Banks, S. Han, K. I. Jang, J. W. Lee, K. T. Lee, X. Feng, Y. Huang, M. Fabiani, G. Gratton, U. Paik, and J. A. Rogers, “Miniaturized Battery-Free Wireless Systems for Wearable Pulse Oximetry,” Adv. Funct. Mater. 27(1), 1604373 (2017). [CrossRef]  

211. H. Li, C. Zhang, and X. Feng, “Monte Carlo simulation of light scattering in tissue for the design of skin-like optical devices,” Biomed. Opt. Express 10(2), 868–878 (2019). [CrossRef]  

212. H. Li, Y. Xu, X. Li, Y. Chen, Y. Jiang, C. Zhang, B. Lu, J. Wang, Y. Ma, Y. Chen, Y. Huang, M. Ding, H. Su, G. Song, Y. Luo, and X. Feng, “Epidermal Inorganic Optoelectronics for Blood Oxygen Measurement,” Adv. Healthcare Mater. 6(9), 1601013 (2017). [CrossRef]  

213. H. Lee, H. Ko, and J. Lee, “Reflectance pulse oximetry: Practical issues and limitations,” ICT Express 2(4), 195–198 (2016). [CrossRef]  

214. D. B. Wax, P. Rubin, and S. Neustein, “A comparison of transmittance and reflectance pulse oximetry during vascular surgery,” Anesth. Analg. 109(6), 1847–1849 (2009). [CrossRef]  

215. S. Sugino, N. Kanaya, M. Mizuuchi, M. Nakayama, and A. Namiki, “Forehead is as sensitive as finger pulse oximetry during general anesthesia,” Can. J. Anaesth. 51(5), 432–436 (2004). [CrossRef]  

216. S. J. Choi, H. J. Ahn, M. K. Yang, C. S. Kim, W. S. Sim, J. A. Kim, J. G. Kang, J. K. Kim, and J. Y. Kang, “Comparison of desaturation and resaturation response times between transmission and reflectance pulse oximeters,” Acta Anaesthesiol. Scand. 54(2), 212–217 (2010). [CrossRef]  

217. S. F. Babikir and R. A. Ismail, “Oxygen Level Measurement Techniques: Pulse Oximetry,” Journal of Science and Technology16(2), (2015).

218. J. A. Rogers, T. Someya, and Y. Huang, “Materials and mechanics for stretchable electronics,” science 327(5973), 1603–1607 (2010). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (10)

Fig. 1.
Fig. 1. Reviews of flexible and stretchable electronics, including flexible LEDs [64]; epidermal electronics [78]; transfer printing technologies [77]; flexible and stretchable bio-electronic devices integrated with nanomaterials [7981]; Copyright 2016, Wiley, Copyright 2008, 2014, Springer Nature; flexible and stretchable electronics [8287]; flexible and stretchable displays [88]; Flexible Electronics; wearable biosensors for healthcare monitoring [89100]. Figures reused with permission from the corresponding publisher.
Fig. 2.
Fig. 2. Mechanical designs of the flexible and stretchable inorganic semiconductors. (a) SEM image of the flexible single crystal ribbon, which thickness is less than 1 µm [105]. (b) Scheme for explaining neutral layer, and finite element simulated through-thickness [106]. (c) Schematic illustration of the process for fabricating wavy ribbons on a plastic substrate [107]. (d) SEM image of GaAs wavy ribbons on PDMS substrate [108]. (e) Some examples of wavy ribbons based on different materials. From top to bottom: Silicon, AlN, PZT, respectively [111111]. (f) Fractal geometries as general layouts for stretchable electronics and FEM images of each structure under elastic tensile strain and the demonstration of Si NMs patterned into Peano layouts then bonded to a 40% pre-strained elastomeric substrate [112]. Figures reused with permission from the corresponding publisher.
Fig. 3.
Fig. 3. Photonic analysis for observing semiconductor properties with different stain. (a) Experiments and results of CL measurements in a bent ZnO NWs. Line-scanning CL spectra on the strain-free section “I” and from the pure bending region (section “II”-“IV”) with constant strain-gradient g = 1.20% µm−1 at 5.5 K, and the three-dimensional graph of the NBE emission spectra [115]. Copyright 2014, Wiley. (b) Optical microscope image, 3D microstructure, and profile details of a single ribbon for three samples with different waveform [116]. The right figures show the PL spectra with continuous position intervals in a single period, normalized Raman spectra of GaAs long-wavelength optical phonon on three particular positions, and periodic band gap shift in one more period of different samples. Copyright 2016, ACS.
Fig. 4.
Fig. 4. Transfer printing technologies. (a) Schematic illustration of the generic process flow for a typical transfer printing process using PDMS stamp [151]. (b) Schematic diagram of critical energy release rates for the ink/substrate interface and the stamp/ink interface. The intersection of the horizontal line in the middle with the curve represents the critical peel velocity for the kinetically controlled transfer printing. The horizontal lines at the bottom and top represent very weak and strong film/substrate interfaces, respectively, corresponding to pick up only and printing only [152]. (c) Schematic illustration of the surface-treatment transfer printing process to pattern the PDMS substrate in the subtractive and addictive switchable transfer modes. The PDMS can be patterned to different shapes, as shown in right optical images [153]. (d) Octopus-inspired pads for the transfer printing process. The adhesive force is reversibly switched by the hydrogel actuation in response to the environmental temperature. The smart stamps can transfer print semiconductors with different materials, such as Si ribbons, InGaAs ribbons, or heterostructure [154]. (e) Programmable transfer printing process via automated laser writing on a micropatterned shape memory polymer stamp. The silicon and GaAs nanoribbons can be hybrid printed on PDMS, and also printed patterns of a cross [155]. Figures reused with permission from the corresponding publisher.
Fig. 5.
Fig. 5. Flexible and stretchable light-emitting devices. SEMs and optical images of the red (a) and (b) blue LED array. The red LEDs are based on AlInGaP structures (50 mm by 50 mm) created by vertical, patterned etching an epitaxial multilayer stack grown on a GaAs wafer [156]. The blue LEDs are InGaN µ-ILED arrays grown on Si (111) wafer [157]. Both of them are moved to a plastic substrate by transfer printing techniques. (c) Wearable micro light-emitting diodes integrated on a lab coat [158]. (d) RGB displays with transfer-printed assemblies of 8 µm × 15 µm inorganic light-emitting diodes [159]. (e) A stretchable high-density array of blue LEDs. There are 500*500 units connected by stretchable wire. Figures reused with permission from the corresponding publisher.
Fig. 6.
Fig. 6. Mechanical designs for the stretchable LED array. (a) Stretchable active-matrix inorganic light-emitting diode display. These images show uniform emission characteristics of the stretchable LED display under uniaxial strain from 0% to 40%; white lines indicate the horizontal direction [173]. (b) Stretchable LED chain with strain isolation design. The substrate can be spatio-programmed rigidity, which means the substrate is soft and stretchable expect the section of LEDs [174]. Figures reused with permission from the corresponding publisher.
Fig. 7.
Fig. 7. Flexible and stretchable inorganic/inorganic-organic hybrid photodetectors. (a) Flexible phototransistors based on transferrable single-crystalline Si nanomembrane (Si NM). The device exhibits stable responsivity with less than 5% of variation under bending at small radii of curvatures (up to 15 mm) [176]. (b) Optical images of the stretchable GaAs photovoltaic modules. The photodetector array is designed as an island-bridge structure for stretchability [177]. (c) Images and switching behaviors of an intrinsically stretchable nanowire photodetector based on ZnO NM. The ZnO NM layer is embedded in the PDMS matrix, yielding a structure with all the components (electrodes and detection channels) being intrinsically stretchable [191]. (d) Silicon-organic hybrid solar cells by the vertical array of Si micro-pillars embedded into elastomeric substrates [192]. (e) Flexibility organic-inorganic perovskite solar cells. The recoverable shape polymer is utilized as a substrate of perovskite solar cell to enable complete shape recovery of the device upon sub-millimeter bending radii [193]. Figures reused with permission from the corresponding publisher.
Fig. 8.
Fig. 8. flexible and stretchable photonics for interconnection and sensing. (a) Integration photonics for signal transmission. The device integrates waveguides, micro-mirrors, and optoelectronic components [202]. (b) Flexible silicon photonic devices. An optical strain sensor uses 5 µm-radius micro ring resonators on the flexible substrate [203]. (c) Single-mode stretchable integrated photonic devices using chalcogenide glass (ChG) and epoxy polymer [204]. (d) Continuously tunable color filter based on a self-assembled isotopically stretchable microbead monolayer. The spectra of the filtered light are solely controlled by external strain (up to 32% radial strain) to cover a broad visible spectrum [205]. Figures reused with permission from the corresponding publisher.
Fig. 9.
Fig. 9. Representative flexible and stretchable inorganic optoelectronics for biomedical applications. (a) Images of components for a digital camera that takes the form of a hemispherical, apposition compound eye. The photodiode array is fabricated by silicon, and the small hemisphere lens is made by PDMS [206]. (b) Geometric origami of silicon optoelectronics for the hemispherical electronic eye. 676 polygon blocks consisting of pentagons and hexagons were mapped into a net of a subdivided half truncated icosahedron, which was then folded to form a hemisphere [207]. (c) Images and performance of the released filter membrane on a flexible PDMS sheet. This filter structure consists of a series of titanium dioxide (TiO2) and silicon dioxide (SiO2) films with a total thickness of about 8 µm. The films are biocompatible, and no apparent signs of pathological inflammatory tissue responses to the filter implantation are observed after 5 weeks [208]. (d) Injectable optoelectronics with applications for wireless optogenetics. A custom flexible polyimide film–based lightweight (∼0.7 g) power scavenger or a rigid printed circuit board-based scavenger (∼2.0 g) can be temporarily mounted on freely moving animals for short-term experimentation without constraint in natural animal behavior [209]. Figures reused with permission from the corresponding publisher.
Fig. 10.
Fig. 10. Flexible and stretchable oximeters. (a) Miniaturized battery-free wireless systems for wearable pulse oximetry. The wireless transmitting components are integrated with inorganic optoelectronics, and the systems can conformal contact to the target probing area (i.e., fingernail, toenail, or various skin locations) for avoiding mechanical damage [210]. (b) The performance analysis via Monte Carlo (MC) simulation. The deformation of flexible devices will cause an unstable optical path. Mechanical optimization is needed to keep the detected signal stable during deformation [211]. (c) Epidermal optoelectronics with strain isolation design for blood oxygen monitoring. The all-in-one free-floating structure is adopted to keep the stable position of LEDs and photodetectors, which stabilize the optical signals for precise blood oxygen monitoring during deformation [212]. Figures reused with permission from the corresponding publisher.

Equations (5)

Equations on this page are rendered with MathJax. Learn more.

E I = E b h 3 12
z ( ε = 0 ) = i = 1 n E I d i [ ( j = 1 i d j ) d i 2 ] / i = 1 n E i d i
λ 0 = π h ε 0 , A 0 = h ε p r e ε c 1
G c s t a m p i n k > G c i n k d o n o r
G c s t a m p i n k < G c i n k r e c e i v e r
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.